13.07.2015 Views

Mathematical Methods for Physics and Engineering - Matematica.NET

Mathematical Methods for Physics and Engineering - Matematica.NET

Mathematical Methods for Physics and Engineering - Matematica.NET

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

This page intentionally left blank


<strong>Mathematical</strong> <strong>Methods</strong> <strong>for</strong> <strong>Physics</strong> <strong>and</strong> <strong>Engineering</strong>The third edition of this highly acclaimed undergraduate textbook is suitable<strong>for</strong> teaching all the mathematics ever likely to be needed <strong>for</strong> an undergraduatecourse in any of the physical sciences. As well as lucid descriptions of all thetopics covered <strong>and</strong> many worked examples, it contains more than 800 exercises.A number of additional topics have been included <strong>and</strong> the text has undergonesignificant reorganisation in some areas. New st<strong>and</strong>-alone chapters:• give a systematic account of the ‘special functions’ of physical science• cover an extended range of practical applications of complex variables includingWKB methods <strong>and</strong> saddle-point integration techniques• provide an introduction to quantum operators.Further tabulations, of relevance in statistics <strong>and</strong> numerical integration, havebeen added. In this edition, all 400 odd-numbered exercises are provided withcomplete worked solutions in a separate manual, available to both students <strong>and</strong>their teachers; these are in addition to the hints <strong>and</strong> outline answers given inthe main text. The even-numbered exercises have no hints, answers or workedsolutions <strong>and</strong> can be used <strong>for</strong> unaided homework; full solutions to them areavailable to instructors on a password-protected website.Ken Riley read mathematics at the University of Cambridge <strong>and</strong> proceededto a Ph.D. there in theoretical <strong>and</strong> experimental nuclear physics. He became aresearch associate in elementary particle physics at Brookhaven, <strong>and</strong> then, havingtaken up a lectureship at the Cavendish Laboratory, Cambridge, continued thisresearch at the Ruther<strong>for</strong>d Laboratory <strong>and</strong> Stan<strong>for</strong>d; in particular he was involvedin the experimental discovery of a number of the early baryonic resonances. Aswell as having been Senior Tutor at Clare College, where he has taught physics<strong>and</strong> mathematics <strong>for</strong> over 40 years, he has served on many committees concernedwith the teaching <strong>and</strong> examining of these subjects at all levels of tertiary <strong>and</strong>undergraduate education. He is also one of the authors of 200 Puzzling <strong>Physics</strong>Problems.Michael Hobson read natural sciences at the University of Cambridge, specialisingin theoretical physics, <strong>and</strong> remained at the Cavendish Laboratory tocomplete a Ph.D. in the physics of star-<strong>for</strong>mation. As a research fellow at TrinityHall, Cambridge <strong>and</strong> subsequently an advanced fellow of the Particle <strong>Physics</strong><strong>and</strong> Astronomy Research Council, he developed an interest in cosmology, <strong>and</strong>in particular in the study of fluctuations in the cosmic microwave background.He was involved in the first detection of these fluctuations using a ground-basedinterferometer. He is currently a University Reader at the Cavendish Laboratory,his research interests include both theoretical <strong>and</strong> observational aspects of cosmology,<strong>and</strong> he is the principal author of General Relativity: An Introduction <strong>for</strong>


Physicists. He is also a Director of Studies in Natural Sciences at Trinity Hall <strong>and</strong>enjoys an active role in the teaching of undergraduate physics <strong>and</strong> mathematics.Stephen Bence obtained both his undergraduate degree in Natural Sciences<strong>and</strong> his Ph.D. in Astrophysics from the University of Cambridge. He then becamea Research Associate with a special interest in star-<strong>for</strong>mation processes <strong>and</strong> thestructure of star-<strong>for</strong>ming regions. In particular, his research concentrated on thephysics of jets <strong>and</strong> outflows from young stars. He has had considerable experienceof teaching mathematics <strong>and</strong> physics to undergraduate <strong>and</strong> pre-universtiystudents.ii


<strong>Mathematical</strong> <strong>Methods</strong><strong>for</strong> <strong>Physics</strong> <strong>and</strong> <strong>Engineering</strong>Third EditionK.F. RILEY, M.P. HOBSON <strong>and</strong> S.J. BENCE


cambridge university pressCambridge, New York, Melbourne, Madrid, Cape Town, Singapore, São PauloCambridge University PressThe Edinburgh Building, Cambridge cb2 2ru, UKPublished in the United States of America by Cambridge University Press, New Yorkwww.cambridge.orgIn<strong>for</strong>mation on this title: www.cambridge.org/9780521861533© K. F. Riley, M. P. Hobson <strong>and</strong> S. J. Bence 2006This publication is in copyright. Subject to statutory exception <strong>and</strong> to the provision ofrelevant collective licensing agreements, no reproduction of any part may take placewithout the written permission of Cambridge University Press.First published in print <strong>for</strong>mat 2006isbn-13 978-0-511-16842-0 eBook (EBL)isbn-10 0-511-16842-x eBook (EBL)isbn-13 978-0-521-86153-3 hardbackisbn-10 0-521-86153-5 hardbackisbn-13 978-0-521-67971-8 paperbackisbn-10 0-521-67971-0 paperbackCambridge University Press has no responsibility <strong>for</strong> the persistence or accuracy of urls<strong>for</strong> external or third-party internet websites referred to in this publication, <strong>and</strong> does notguarantee that any content on such websites is, or will remain, accurate or appropriate.


ContentsPreface to the third editionPreface to the second editionPreface to the first editionpage xxxxiiixxv1 Preliminary algebra 11.1 Simple functions <strong>and</strong> equations 1Polynomial equations; factorisation; properties of roots1.2 Trigonometric identities 10Single angle; compound angles; double- <strong>and</strong> half-angle identities1.3 Coordinate geometry 151.4 Partial fractions 18Complications <strong>and</strong> special cases1.5 Binomial expansion 251.6 Properties of binomial coefficients 271.7 Some particular methods of proof 30Proof by induction; proof by contradiction; necessary <strong>and</strong> sufficient conditions1.8 Exercises 361.9 Hints <strong>and</strong> answers 392 Preliminary calculus 412.1 Differentiation 41Differentiation from first principles; products; the chain rule; quotients;implicit differentiation; logarithmic differentiation; Leibnitz’ theorem; specialpoints of a function; curvature; theorems of differentiationv


CONTENTS2.2 Integration 59Integration from first principles; the inverse of differentiation; by inspection;sinusoidal functions; logarithmic integration; using partial fractions;substitution method; integration by parts; reduction <strong>for</strong>mulae; infinite <strong>and</strong>improper integrals; plane polar coordinates; integral inequalities; applicationsof integration2.3 Exercises 762.4 Hints <strong>and</strong> answers 813 Complex numbers <strong>and</strong> hyperbolic functions 833.1 The need <strong>for</strong> complex numbers 833.2 Manipulation of complex numbers 85Addition <strong>and</strong> subtraction; modulus <strong>and</strong> argument; multiplication; complexconjugate; division3.3 Polar representation of complex numbers 92Multiplication <strong>and</strong> division in polar <strong>for</strong>m3.4 de Moivre’s theorem 95trigonometric identities; finding the nth roots of unity; solving polynomialequations3.5 Complex logarithms <strong>and</strong> complex powers 993.6 Applications to differentiation <strong>and</strong> integration 1013.7 Hyperbolic functions 102Definitions; hyperbolic–trigonometric analogies; identities of hyperbolicfunctions; solving hyperbolic equations; inverses of hyperbolic functions;calculus of hyperbolic functions3.8 Exercises 1093.9 Hints <strong>and</strong> answers 1134 Series <strong>and</strong> limits 1154.1 Series 1154.2 Summation of series 116Arithmetic series; geometric series; arithmetico-geometric series; the differencemethod; series involving natural numbers; trans<strong>for</strong>mation of series4.3 Convergence of infinite series 124Absolute <strong>and</strong> conditional convergence; series containing only real positiveterms; alternating series test4.4 Operations with series 1314.5 Power series 131Convergence of power series; operations with power series4.6 Taylor series 136Taylor’s theorem; approximation errors; st<strong>and</strong>ard Maclaurin series4.7 Evaluation of limits 1414.8 Exercises 1444.9 Hints <strong>and</strong> answers 149vi


CONTENTS5 Partial differentiation 1515.1 Definition of the partial derivative 1515.2 The total differential <strong>and</strong> total derivative 1535.3 Exact <strong>and</strong> inexact differentials 1555.4 Useful theorems of partial differentiation 1575.5 The chain rule 1575.6 Change of variables 1585.7 Taylor’s theorem <strong>for</strong> many-variable functions 1605.8 Stationary values of many-variable functions 1625.9 Stationary values under constraints 1675.10 Envelopes 1735.11 Thermodynamic relations 1765.12 Differentiation of integrals 1785.13 Exercises 1795.14 Hints <strong>and</strong> answers 1856 Multiple integrals 1876.1 Double integrals 1876.2 Triple integrals 1906.3 Applications of multiple integrals 191Areas <strong>and</strong> volumes; masses, centres of mass <strong>and</strong> centroids; Pappus’ theorems;moments of inertia; mean values of functions6.4 Change of variables in multiple integrals 199Change∫of variables in double integrals; evaluation of the integral I =∞−∞ e−x2 dx; change of variables in triple integrals; general properties ofJacobians6.5 Exercises 2076.6 Hints <strong>and</strong> answers 2117 Vector algebra 2127.1 Scalars <strong>and</strong> vectors 2127.2 Addition <strong>and</strong> subtraction of vectors 2137.3 Multiplication by a scalar 2147.4 Basis vectors <strong>and</strong> components 2177.5 Magnitude of a vector 2187.6 Multiplication of vectors 219Scalar product; vector product; scalar triple product; vector triple productvii


CONTENTS7.7 Equations of lines, planes <strong>and</strong> spheres 2267.8 Using vectors to find distances 229Point to line; point to plane; line to line; line to plane7.9 Reciprocal vectors 2337.10 Exercises 2347.11 Hints <strong>and</strong> answers 2408 Matrices <strong>and</strong> vector spaces 2418.1 Vector spaces 242Basis vectors; inner product; some useful inequalities8.2 Linear operators 2478.3 Matrices 2498.4 Basic matrix algebra 250Matrix addition; multiplication by a scalar; matrix multiplication8.5 Functions of matrices 2558.6 The transpose of a matrix 2558.7 The complex <strong>and</strong> Hermitian conjugates of a matrix 2568.8 The trace of a matrix 2588.9 The determinant of a matrix 259Properties of determinants8.10 The inverse of a matrix 2638.11 The rank of a matrix 2678.12 Special types of square matrix 268Diagonal; triangular; symmetric <strong>and</strong> antisymmetric; orthogonal; Hermitian<strong>and</strong> anti-Hermitian; unitary; normal8.13 Eigenvectors <strong>and</strong> eigenvalues 272Of a normal matrix; of Hermitian <strong>and</strong> anti-Hermitian matrices; of a unitarymatrix; of a general square matrix8.14 Determination of eigenvalues <strong>and</strong> eigenvectors 280Degenerate eigenvalues8.15 Change of basis <strong>and</strong> similarity trans<strong>for</strong>mations 2828.16 Diagonalisation of matrices 2858.17 Quadratic <strong>and</strong> Hermitian <strong>for</strong>ms 288Stationary properties of the eigenvectors; quadratic surfaces8.18 Simultaneous linear equations 292Range; null space; N simultaneous linear equations in N unknowns; singularvalue decomposition8.19 Exercises 3078.20 Hints <strong>and</strong> answers 3149 Normal modes 3169.1 Typical oscillatory systems 3179.2 Symmetry <strong>and</strong> normal modes 322viii


CONTENTS9.3 Rayleigh–Ritz method 3279.4 Exercises 3299.5 Hints <strong>and</strong> answers 33210 Vector calculus 33410.1 Differentiation of vectors 334Composite vector expressions; differential of a vector10.2 Integration of vectors 33910.3 Space curves 34010.4 Vector functions of several arguments 34410.5 Surfaces 34510.6 Scalar <strong>and</strong> vector fields 34710.7 Vector operators 347Gradient of a scalar field; divergence of a vector field; curl of a vector field10.8 Vector operator <strong>for</strong>mulae 354Vector operators acting on sums <strong>and</strong> products; combinations of grad, div <strong>and</strong>curl10.9 Cylindrical <strong>and</strong> spherical polar coordinates 35710.10 General curvilinear coordinates 36410.11 Exercises 36910.12 Hints <strong>and</strong> answers 37511 Line, surface <strong>and</strong> volume integrals 37711.1 Line integrals 377Evaluating line integrals; physical examples; line integrals with respect to ascalar11.2 Connectivity of regions 38311.3 Green’s theorem in a plane 38411.4 Conservative fields <strong>and</strong> potentials 38711.5 Surface integrals 389Evaluating surface integrals; vector areas of surfaces; physical examples11.6 Volume integrals 396Volumes of three-dimensional regions11.7 Integral <strong>for</strong>ms <strong>for</strong> grad, div <strong>and</strong> curl 39811.8 Divergence theorem <strong>and</strong> related theorems 401Green’s theorems; other related integral theorems; physical applications11.9 Stokes’ theorem <strong>and</strong> related theorems 406Related integral theorems; physical applications11.10 Exercises 40911.11 Hints <strong>and</strong> answers 41412 Fourier series 41512.1 The Dirichlet conditions 415ix


CONTENTS12.2 The Fourier coefficients 41712.3 Symmetry considerations 41912.4 Discontinuous functions 42012.5 Non-periodic functions 42212.6 Integration <strong>and</strong> differentiation 42412.7 Complex Fourier series 42412.8 Parseval’s theorem 42612.9 Exercises 42712.10 Hints <strong>and</strong> answers 43113 Integral trans<strong>for</strong>ms 43313.1 Fourier trans<strong>for</strong>ms 433The uncertainty principle; Fraunhofer diffraction; the Dirac δ-function;relation of the δ-function to Fourier trans<strong>for</strong>ms; properties of Fouriertrans<strong>for</strong>ms; odd <strong>and</strong> even functions; convolution <strong>and</strong> deconvolution; correlationfunctions <strong>and</strong> energy spectra; Parseval’s theorem; Fourier trans<strong>for</strong>ms in higherdimensions13.2 Laplace trans<strong>for</strong>ms 453Laplace trans<strong>for</strong>ms of derivatives <strong>and</strong> integrals; other properties of Laplacetrans<strong>for</strong>ms13.3 Concluding remarks 45913.4 Exercises 46013.5 Hints <strong>and</strong> answers 46614 First-order ordinary differential equations 46814.1 General <strong>for</strong>m of solution 46914.2 First-degree first-order equations 470Separable-variable equations; exact equations; inexact equations, integratingfactors; linear equations; homogeneous equations; isobaric equations;Bernoulli’s equation; miscellaneous equations14.3 Higher-degree first-order equations 480Equations soluble <strong>for</strong> p; <strong>for</strong>x; <strong>for</strong>y; Clairaut’s equation14.4 Exercises 48414.5 Hints <strong>and</strong> answers 48815 Higher-order ordinary differential equations 49015.1 Linear equations with constant coefficients 492Finding the complementary function y c (x); finding the particular integraly p (x); constructing the general solution y c (x) +y p (x); linear recurrencerelations; Laplace trans<strong>for</strong>m method15.2 Linear equations with variable coefficients 503The Legendre <strong>and</strong> Euler linear equations; exact equations; partially knowncomplementary function; variation of parameters; Green’s functions; canonical<strong>for</strong>m <strong>for</strong> second-order equationsx


CONTENTS15.3 General ordinary differential equations 518Dependent variable absent; independent variable absent; non-linear exactequations; isobaric or homogeneous equations; equations homogeneous in xor y alone; equations having y = Ae x as a solution15.4 Exercises 52315.5 Hints <strong>and</strong> answers 52916 Series solutions of ordinary differential equations 53116.1 Second-order linear ordinary differential equations 531Ordinary <strong>and</strong> singular points16.2 Series solutions about an ordinary point 53516.3 Series solutions about a regular singular point 538Distinct roots not differing by an integer; repeated root of the indicialequation; distinct roots differing by an integer16.4 Obtaining a second solution 544The Wronskian method; the derivative method; series <strong>for</strong>m of the secondsolution16.5 Polynomial solutions 54816.6 Exercises 55016.7 Hints <strong>and</strong> answers 55317 Eigenfunction methods <strong>for</strong> differential equations 55417.1 Sets of functions 556Some useful inequalities17.2 Adjoint, self-adjoint <strong>and</strong> Hermitian operators 55917.3 Properties of Hermitian operators 561Reality of the eigenvalues; orthogonality of the eigenfunctions; constructionof real eigenfunctions17.4 Sturm–Liouville equations 564Valid boundary conditions; putting an equation into Sturm–Liouville <strong>for</strong>m17.5 Superposition of eigenfunctions: Green’s functions 56917.6 A useful generalisation 57217.7 Exercises 57317.8 Hints <strong>and</strong> answers 57618 Special functions 57718.1 Legendre functions 577General solution <strong>for</strong> integer l; properties of Legendre polynomials18.2 Associated Legendre functions 58718.3 Spherical harmonics 59318.4 Chebyshev functions 59518.5 Bessel functions 602General solution <strong>for</strong> non-integer ν; general solution <strong>for</strong> integer ν; propertiesof Bessel functionsxi


CONTENTS18.6 Spherical Bessel functions 61418.7 Laguerre functions 61618.8 Associated Laguerre functions 62118.9 Hermite functions 62418.10 Hypergeometric functions 62818.11 Confluent hypergeometric functions 63318.12 The gamma function <strong>and</strong> related functions 63518.13 Exercises 64018.14 Hints <strong>and</strong> answers 64619 Quantum operators 64819.1 Operator <strong>for</strong>malism 648Commutators19.2 Physical examples of operators 656Uncertainty principle; angular momentum; creation <strong>and</strong> annihilation operators19.3 Exercises 67119.4 Hints <strong>and</strong> answers 67420 Partial differential equations: general <strong>and</strong> particular solutions 67520.1 Important partial differential equations 676The wave equation; the diffusion equation; Laplace’s equation; Poisson’sequation; Schrödinger’s equation20.2 General <strong>for</strong>m of solution 68020.3 General <strong>and</strong> particular solutions 681First-order equations; inhomogeneous equations <strong>and</strong> problems; second-orderequations20.4 The wave equation 69320.5 The diffusion equation 69520.6 Characteristics <strong>and</strong> the existence of solutions 699First-order equations; second-order equations20.7 Uniqueness of solutions 70520.8 Exercises 70720.9 Hints <strong>and</strong> answers 71121 Partial differential equations: separation of variables<strong>and</strong> other methods 71321.1 Separation of variables: the general method 71321.2 Superposition of separated solutions 71721.3 Separation of variables in polar coordinates 725Laplace’s equation in polar coordinates; spherical harmonics; other equationsin polar coordinates; solution by expansion; separation of variables <strong>for</strong>inhomogeneous equations21.4 Integral trans<strong>for</strong>m methods 747xii


CONTENTS21.5 Inhomogeneous problems – Green’s functions 751Similarities to Green’s functions <strong>for</strong> ordinary differential equations; generalboundary-value problems; Dirichlet problems; Neumann problems21.6 Exercises 76721.7 Hints <strong>and</strong> answers 77322 Calculus of variations 77522.1 The Euler–Lagrange equation 77622.2 Special cases 777F does not contain y explicitly; F does not contain x explicitly22.3 Some extensions 781Several dependent variables; several independent variables; higher-orderderivatives; variable end-points22.4 Constrained variation 78522.5 Physical variational principles 787Fermat’s principle in optics; Hamilton’s principle in mechanics22.6 General eigenvalue problems 79022.7 Estimation of eigenvalues <strong>and</strong> eigenfunctions 79222.8 Adjustment of parameters 79522.9 Exercises 79722.10 Hints <strong>and</strong> answers 80123 Integral equations 80323.1 Obtaining an integral equation from a differential equation 80323.2 Types of integral equation 80423.3 Operator notation <strong>and</strong> the existence of solutions 80523.4 Closed-<strong>for</strong>m solutions 806Separable kernels; integral trans<strong>for</strong>m methods; differentiation23.5 Neumann series 81323.6 Fredholm theory 81523.7 Schmidt–Hilbert theory 81623.8 Exercises 81923.9 Hints <strong>and</strong> answers 82324 Complex variables 82424.1 Functions of a complex variable 82524.2 The Cauchy–Riemann relations 82724.3 Power series in a complex variable 83024.4 Some elementary functions 83224.5 Multivalued functions <strong>and</strong> branch cuts 83524.6 Singularities <strong>and</strong> zeros of complex functions 83724.7 Con<strong>for</strong>mal trans<strong>for</strong>mations 83924.8 Complex integrals 845xiii


CONTENTS24.9 Cauchy’s theorem 84924.10 Cauchy’s integral <strong>for</strong>mula 85124.11 Taylor <strong>and</strong> Laurent series 85324.12 Residue theorem 85824.13 Definite integrals using contour integration 86124.14 Exercises 86724.15 Hints <strong>and</strong> answers 87025 Applications of complex variables 87125.1 Complex potentials 87125.2 Applications of con<strong>for</strong>mal trans<strong>for</strong>mations 87625.3 Location of zeros 87925.4 Summation of series 88225.5 Inverse Laplace trans<strong>for</strong>m 88425.6 Stokes’ equation <strong>and</strong> Airy integrals 88825.7 WKB methods 89525.8 Approximations to integrals 905Level lines <strong>and</strong> saddle points; steepest descents; stationary phase25.9 Exercises 92025.10 Hints <strong>and</strong> answers 92526 Tensors 92726.1 Some notation 92826.2 Change of basis 92926.3 Cartesian tensors 93026.4 First- <strong>and</strong> zero-order Cartesian tensors 93226.5 Second- <strong>and</strong> higher-order Cartesian tensors 93526.6 The algebra of tensors 93826.7 The quotient law 93926.8 The tensors δ ij <strong>and</strong> ɛ ijk 94126.9 Isotropic tensors 94426.10 Improper rotations <strong>and</strong> pseudotensors 94626.11 Dual tensors 94926.12 Physical applications of tensors 95026.13 Integral theorems <strong>for</strong> tensors 95426.14 Non-Cartesian coordinates 95526.15 The metric tensor 95726.16 General coordinate trans<strong>for</strong>mations <strong>and</strong> tensors 96026.17 Relative tensors 96326.18 Derivatives of basis vectors <strong>and</strong> Christoffel symbols 96526.19 Covariant differentiation 96826.20 Vector operators in tensor <strong>for</strong>m 971xiv


CONTENTS26.21 Absolute derivatives along curves 97526.22 Geodesics 97626.23 Exercises 97726.24 Hints <strong>and</strong> answers 98227 Numerical methods 98427.1 Algebraic <strong>and</strong> transcendental equations 985Rearrangement of the equation; linear interpolation; binary chopping;Newton–Raphson method27.2 Convergence of iteration schemes 99227.3 Simultaneous linear equations 994Gaussian elimination; Gauss–Seidel iteration; tridiagonal matrices27.4 Numerical integration 1000Trapezium rule; Simpson’s rule; Gaussian integration; Monte Carlo methods27.5 Finite differences 101927.6 Differential equations 1020Difference equations; Taylor series solutions; prediction <strong>and</strong> correction;Runge–Kutta methods; isoclines27.7 Higher-order equations 102827.8 Partial differential equations 103027.9 Exercises 103327.10 Hints <strong>and</strong> answers 103928 Group theory 104128.1 Groups 1041Definition of a group; examples of groups28.2 Finite groups 104928.3 Non-Abelian groups 105228.4 Permutation groups 105628.5 Mappings between groups 105928.6 Subgroups 106128.7 Subdividing a group 1063Equivalence relations <strong>and</strong> classes; congruence <strong>and</strong> cosets; conjugates <strong>and</strong>classes28.8 Exercises 107028.9 Hints <strong>and</strong> answers 107429 Representation theory 107629.1 Dipole moments of molecules 107729.2 Choosing an appropriate <strong>for</strong>malism 107829.3 Equivalent representations 108429.4 Reducibility of a representation 108629.5 The orthogonality theorem <strong>for</strong> irreducible representations 1090xv


CONTENTS29.6 Characters 1092Orthogonality property of characters29.7 Counting irreps using characters 1095Summation rules <strong>for</strong> irreps29.8 Construction of a character table 110029.9 Group nomenclature 110229.10 Product representations 110329.11 Physical applications of group theory 1105Bonding in molecules; matrix elements in quantum mechanics; degeneracy ofnormal modes; breaking of degeneracies29.12 Exercises 111329.13 Hints <strong>and</strong> answers 111730 Probability 111930.1 Venn diagrams 111930.2 Probability 1124Axioms <strong>and</strong> theorems; conditional probability; Bayes’ theorem30.3 Permutations <strong>and</strong> combinations 113330.4 R<strong>and</strong>om variables <strong>and</strong> distributions 1139Discrete r<strong>and</strong>om variables; continuous r<strong>and</strong>om variables30.5 Properties of distributions 1143Mean; mode <strong>and</strong> median; variance <strong>and</strong> st<strong>and</strong>ard deviation; moments; centralmoments30.6 Functions of r<strong>and</strong>om variables 115030.7 Generating functions 1157Probability generating functions; moment generating functions; characteristicfunctions; cumulant generating functions30.8 Important discrete distributions 1168Binomial; geometric; negative binomial; hypergeometric; Poisson30.9 Important continuous distributions 1179Gaussian; log-normal; exponential; gamma; chi-squared; Cauchy; Breit–Wigner; uni<strong>for</strong>m30.10 The central limit theorem 119530.11 Joint distributions 1196Discrete bivariate; continuous bivariate; marginal <strong>and</strong> conditional distributions30.12 Properties of joint distributions 1199Means; variances; covariance <strong>and</strong> correlation30.13 Generating functions <strong>for</strong> joint distributions 120530.14 Trans<strong>for</strong>mation of variables in joint distributions 120630.15 Important joint distributions 1207Multinominal; multivariate Gaussian30.16 Exercises 121130.17 Hints <strong>and</strong> answers 1219xvi


CONTENTS31 Statistics 122131.1 Experiments, samples <strong>and</strong> populations 122131.2 Sample statistics 1222Averages; variance <strong>and</strong> st<strong>and</strong>ard deviation; moments; covariance <strong>and</strong> correlation31.3 Estimators <strong>and</strong> sampling distributions 1229Consistency, bias <strong>and</strong> efficiency; Fisher’s inequality; st<strong>and</strong>ard errors; confidencelimits31.4 Some basic estimators 1243Mean; variance; st<strong>and</strong>ard deviation; moments; covariance <strong>and</strong> correlation31.5 Maximum-likelihood method 1255ML estimator; trans<strong>for</strong>mation invariance <strong>and</strong> bias; efficiency; errors <strong>and</strong>confidence limits; Bayesian interpretation; large-N behaviour; extendedML method31.6 The method of least squares 1271Linear least squares; non-linear least squares31.7 Hypothesis testing 1277Simple <strong>and</strong> composite hypotheses; statistical tests; Neyman–Pearson; generalisedlikelihood-ratio; Student’s t; Fisher’sF; goodness of fit31.8 Exercises 129831.9 Hints <strong>and</strong> answers 1303Index 1305xvii


CONTENTSI am the very Model <strong>for</strong> a Student <strong>Mathematical</strong>I am the very model <strong>for</strong> a student mathematical;I’ve in<strong>for</strong>mation rational, <strong>and</strong> logical <strong>and</strong> practical.I know the laws of algebra, <strong>and</strong> find them quite symmetrical,And even know the meaning of ‘a variate antithetical’.I’m extremely well acquainted, with all things mathematical.I underst<strong>and</strong> equations, both the simple <strong>and</strong> quadratical.About binomial theorems I’m teeming with a lot o’news,With many cheerful facts about the square of the hypotenuse.I’m very good at integral <strong>and</strong> differential calculus,And solving paradoxes that so often seem to rankle us.In short in matters rational, <strong>and</strong> logical <strong>and</strong> practical,I am the very model <strong>for</strong> a student mathematical.I know the singularities of equations differential,And some of these are regular, but the rest are quite essential.I quote the results of giants; with Euler, Newton, Gauss, Laplace,And can calculate an orbit, given a centre, <strong>for</strong>ce <strong>and</strong> mass.I can reconstruct equations, both canonical <strong>and</strong> <strong>for</strong>mal,And write all kinds of matrices, orthogonal, real <strong>and</strong> normal.I show how to tackle problems that one has never met be<strong>for</strong>e,By analogy or example, or with some clever metaphor.I seldom use equivalence to help decide upon a class,But often find an integral, using a contour o’er a pass.In short in matters rational, <strong>and</strong> logical <strong>and</strong> practical,I am the very model <strong>for</strong> a student mathematical.When you have learnt just what is meant by ‘Jacobian’ <strong>and</strong> ‘Abelian’;When you at sight can estimate, <strong>for</strong> the modal, mean <strong>and</strong> median;When describing normal subgroups is much more than recitation;When you underst<strong>and</strong> precisely what is ‘quantum excitation’;When you know enough statistics that you can recognise RV;When you have learnt all advances that have been made in SVD;And when you can spot the trans<strong>for</strong>m that solves some tricky PDE,You will feel no better student has ever sat <strong>for</strong> a degree.Your accumulated knowledge, whilst extensive <strong>and</strong> exemplary,Will have only been brought down to the beginning of last century,But still in matters rational, <strong>and</strong> logical <strong>and</strong> practical,You’ll be the very model of a student mathematical.xixKFR, with apologies to W.S. Gilbert


Preface to the third editionAs is natural, in the four years since the publication of the second edition ofthis book we have somewhat modified our views on what should be included<strong>and</strong> how it should be presented. In this new edition, although the range of topicscovered has been extended, there has been no significant shift in the general levelof difficulty or in the degree of mathematical sophistication required. Further, wehave aimed to preserve the same style of presentation as seems to have been wellreceived in the first two editions. However, a significant change has been madeto the <strong>for</strong>mat of the chapters, specifically to the way that the exercises, togetherwith their hints <strong>and</strong> answers, have been treated; the details of the change areexplained below.The two major chapters that are new in this third edition are those dealing with‘special functions’ <strong>and</strong> the applications of complex variables. The <strong>for</strong>mer presentsa systematic account of those functions that appear to have arisen in a moreor less haphazard way as a result of studying particular physical situations, <strong>and</strong>are deemed ‘special’ <strong>for</strong> that reason. The treatment presented here shows that,in fact, they are nearly all particular cases of the hypergeometric or confluenthypergeometric functions, <strong>and</strong> are special only in the sense that the parametersof the relevant function take simple or related values.The second new chapter describes how the properties of complex variables canbe used to tackle problems arising from the description of physical situationsor from other seemingly unrelated areas of mathematics. To topics treated inearlier editions, such as the solution of Laplace’s equation in two dimensions, thesummation of series, the location of zeros of polynomials <strong>and</strong> the calculation ofinverse Laplace trans<strong>for</strong>ms, has been added new material covering Airy integrals,saddle-point methods <strong>for</strong> contour integral evaluation, <strong>and</strong> the WKB approach toasymptotic <strong>for</strong>ms.Other new material includes a st<strong>and</strong>-alone chapter on the use of coordinate-freeoperators to establish valuable results in the field of quantum mechanics; amongstxx


PREFACE TO THE THIRD EDITIONthe physical topics covered are angular momentum <strong>and</strong> uncertainty principles.There are also significant additions to the treatment of numerical integration.In particular, Gaussian quadrature based on Legendre, Laguerre, Hermite <strong>and</strong>Chebyshev polynomials is discussed, <strong>and</strong> appropriate tables of points <strong>and</strong> weightsare provided.We now turn to the most obvious change to the <strong>for</strong>mat of the book, namelythe way that the exercises, hints <strong>and</strong> answers are treated. The second edition of<strong>Mathematical</strong> <strong>Methods</strong> <strong>for</strong> <strong>Physics</strong> <strong>and</strong> <strong>Engineering</strong> carried more than twice asmany exercises, based on its various chapters, as did the first. In its preface wediscussed the general question of how such exercises should be treated but, inthe end, decided to provide hints <strong>and</strong> outline answers to all problems, as in thefirst edition. This decision was an uneasy one as, on the one h<strong>and</strong>, it did notallow the exercises to be set as totally unaided homework that could be used <strong>for</strong>assessment purposes but, on the other, it did not give a full explanation of howto tackle a problem when a student needed explicit guidance or a model answer.In order to allow both of these educationally desirable goals to be achieved,we have, in this third edition, completely changed the way in which this matteris h<strong>and</strong>led. A large number of exercises have been included in the penultimatesubsections of the appropriate, sometimes reorganised, chapters. Hints <strong>and</strong> outlineanswers are given, as previously, in the final subsections, but only <strong>for</strong> the oddnumberedexercises. This leaves all even-numbered exercises free to be set asunaided homework, as described below.For the four hundred plus odd-numbered exercises, complete solutions areavailable, to both students <strong>and</strong> their teachers, in the <strong>for</strong>m of a separate manual,Student Solutions Manual <strong>for</strong> <strong>Mathematical</strong> <strong>Methods</strong> <strong>for</strong> <strong>Physics</strong> <strong>and</strong> <strong>Engineering</strong>(Cambridge: Cambridge University Press, 2006); the hints <strong>and</strong> outline answersgiven in this main text are brief summaries of the model answers given in themanual. There, each original exercise is reproduced <strong>and</strong> followed by a fullyworked solution. For those original exercises that make internal reference to thistext or to other (even-numbered) exercises not included in the solutions manual,the questions have been reworded, usually by including additional in<strong>for</strong>mation,so that the questions can st<strong>and</strong> alone.In many cases, the solution given in the manual is even fuller than one thatmight be expected of a good student that has understood the material. This isbecause we have aimed to make the solutions instructional as well as utilitarian.To this end, we have included comments that are intended to show how theplan <strong>for</strong> the solution is fomulated <strong>and</strong> have given the justifications <strong>for</strong> particularintermediate steps (something not always done, even by the best of students). Wehave also tried to write each individual substituted <strong>for</strong>mula in the <strong>for</strong>m that bestindicates how it was obtained, be<strong>for</strong>e simplifying it at the next or a subsequentstage. Where several lines of algebraic manipulation or calculus are needed toobtain a final result, they are normally included in full; this should enable thexxi


PREFACE TO THE THIRD EDITIONstudent to determine whether an incorrect answer is due to a misunderst<strong>and</strong>ingof principles or to a technical error.The remaining four hundred or so even-numbered exercises have no hints oranswers, outlined or detailed, available <strong>for</strong> general access. They can there<strong>for</strong>e beused by instructors as a basis <strong>for</strong> setting unaided homework. Full solutions tothese exercises, in the same general <strong>for</strong>mat as those appearing in the manual(though they may contain references to the main text or to other exercises), areavailable without charge to accredited teachers as downloadable pdf files on thepassword-protected website http://www.cambridge.org/9780521679718. Teacherswishing to have access to the website should contact solutions@cambridge.org<strong>for</strong> registration details.In all new publications, errors <strong>and</strong> typographical mistakes are virtually unavoidable,<strong>and</strong> we would be grateful to any reader who brings instances toour attention. Retrospectively, we would like to record our thanks to ReinhardGerndt, Paul Renteln <strong>and</strong> Joe Tenn <strong>for</strong> making us aware of some errors inthe second edition. Finally, we are extremely grateful to Dave Green <strong>for</strong> hisconsiderable <strong>and</strong> continuing advice concerning L A TEX.Ken Riley, Michael Hobson,Cambridge, 2006xxii


Preface to the second editionSince the publication of the first edition of this book, both through teaching thematerial it covers <strong>and</strong> as a result of receiving helpful comments from colleagues,we have become aware of the desirability of changes in a number of areas.The most important of these is that the mathematical preparation of currentsenior college <strong>and</strong> university entrants is now less thorough than it used to be.To match this, we decided to include a preliminary chapter covering areas suchas polynomial equations, trigonometric identities, coordinate geometry, partialfractions, binomial expansions, necessary <strong>and</strong> sufficient condition <strong>and</strong> proof byinduction <strong>and</strong> contradiction.Whilst the general level of what is included in this second edition has notbeen raised, some areas have been exp<strong>and</strong>ed to take in topics we now feel werenot adequately covered in the first. In particular, increased attention has beengiven to non-square sets of simultaneous linear equations <strong>and</strong> their associatedmatrices. We hope that this more extended treatment, together with the inclusionof singular value matrix decomposition, will make the material of more practicaluse to engineering students. In the same spirit, an elementary treatment of linearrecurrence relations has been included. The topic of normal modes has been givena small chapter of its own, though the links to matrices on the one h<strong>and</strong>, <strong>and</strong> torepresentation theory on the other, have not been lost.Elsewhere, the presentation of probability <strong>and</strong> statistics has been reorganised togive the two aspects more nearly equal weights. The early part of the probabilitychapter has been rewritten in order to present a more coherent developmentbased on Boolean algebra, the fundamental axioms of probability theory <strong>and</strong>the properties of intersections <strong>and</strong> unions. Whilst this is somewhat more <strong>for</strong>malthan previously, we think that it has not reduced the accessibility of these topics<strong>and</strong> hope that it has increased it. The scope of the chapter has been somewhatextended to include all physically important distributions <strong>and</strong> an introduction tocumulants.xxiii


PREFACE TO THE SECOND EDITIONStatistics now occupies a substantial chapter of its own, one that includes systematicdiscussions of estimators <strong>and</strong> their efficiency, sample distributions <strong>and</strong> t-<strong>and</strong> F-tests <strong>for</strong> comparing means <strong>and</strong> variances. Other new topics are applicationsof the chi-squared distribution, maximum-likelihood parameter estimation <strong>and</strong>least-squares fitting. In other chapters we have added material on the followingtopics: curvature, envelopes, curve-sketching, more refined numerical methods<strong>for</strong> differential equations <strong>and</strong> the elements of integration using Monte Carlotechniques.Over the last four years we have received somewhat mixed feedback aboutthe number of exercises at the ends of the various chapters. After consideration,we decided to increase the number substantially, partly to correspond to theadditional topics covered in the text but mainly to give both students <strong>and</strong>their teachers a wider choice. There are now nearly 800 such exercises, many withseveral parts. An even more vexed question has been whether to provide hints <strong>and</strong>answers to all the exercises or just to ‘the odd-numbered’ ones, as is the normalpractice <strong>for</strong> textbooks in the United States, thus making the remainder moresuitable <strong>for</strong> setting as homework. In the end, we decided that hints <strong>and</strong> outlinesolutions should be provided <strong>for</strong> all the exercises, in order to facilitate independentstudy while leaving the details of the calculation as a task <strong>for</strong> the student.In conclusion, we hope that this edition will be thought by its users to be‘heading in the right direction’ <strong>and</strong> would like to place on record our thanks toall who have helped to bring about the changes <strong>and</strong> adjustments. Naturally, thosecolleagues who have noted errors or ambiguities in the first edition <strong>and</strong> broughtthem to our attention figure high on the list, as do the staff at The CambridgeUniversity Press. In particular, we are grateful to Dave Green <strong>for</strong> continued L A TEXadvice, Susan Parkinson <strong>for</strong> copy-editing the second edition with her usual keeneye <strong>for</strong> detail <strong>and</strong> flair <strong>for</strong> crafting coherent prose <strong>and</strong> Alison Woollatt <strong>for</strong> onceagain turning our basic L A TEX into a beautifully typeset book. Our thanks goto all of them, though of course we accept full responsibility <strong>for</strong> any remainingerrors or ambiguities, of which, as with any new publication, there are bound tobe some.On a more personal note, KFR again wishes to thank his wife Penny <strong>for</strong> herunwavering support, not only in his academic <strong>and</strong> tutorial work, but also in theirjoint ef<strong>for</strong>ts to convert time at the bridge table into ‘green points’ on their record.MPH is once more indebted to his wife, Becky, <strong>and</strong> his mother, Pat, <strong>for</strong> theirtireless support <strong>and</strong> encouragement above <strong>and</strong> beyond the call of duty. MPHdedicates his contribution to this book to the memory of his father, RonaldLeonard Hobson, whose gentle kindness, patient underst<strong>and</strong>ing <strong>and</strong> unbreakablespirit made all things seem possible.Ken Riley, Michael HobsonCambridge, 2002xxiv


Preface to the first editionA knowledge of mathematical methods is important <strong>for</strong> an increasing number ofuniversity <strong>and</strong> college courses, particularly in physics, engineering <strong>and</strong> chemistry,but also in more general science. Students embarking on such courses come fromdiverse mathematical backgrounds, <strong>and</strong> their core knowledge varies considerably.We have there<strong>for</strong>e decided to write a textbook that assumes knowledge only ofmaterial that can be expected to be familiar to all the current generation ofstudents starting physical science courses at university. In the United Kingdomthis corresponds to the st<strong>and</strong>ard of Mathematics A-level, whereas in the UnitedStates the material assumed is that which would normally be covered at juniorcollege.Starting from this level, the first six chapters cover a collection of topicswith which the reader may already be familiar, but which are here extended<strong>and</strong> applied to typical problems encountered by first-year university students.They are aimed at providing a common base of general techniques used inthe development of the remaining chapters. Students who have had additionalpreparation, such as Further Mathematics at A-level, will find much of thismaterial straight<strong>for</strong>ward.Following these opening chapters, the remainder of the book is intended tocover at least that mathematical material which an undergraduate in the physicalsciences might encounter up to the end of his or her course. The book is alsoappropriate <strong>for</strong> those beginning graduate study with a mathematical content, <strong>and</strong>naturally much of the material <strong>for</strong>ms parts of courses <strong>for</strong> mathematics students.Furthermore, the text should provide a useful reference <strong>for</strong> research workers.The general aim of the book is to present a topic in three stages. The firststage is a qualitative introduction, wherever possible from a physical point ofview. The second is a more <strong>for</strong>mal presentation, although we have deliberatelyavoided strictly mathematical questions such as the existence of limits, uni<strong>for</strong>mconvergence, the interchanging of integration <strong>and</strong> summation orders, etc. on thexxv


PREFACE TO THE FIRST EDITIONgrounds that ‘this is the real world; it must behave reasonably’. Finally a workedexample is presented, often drawn from familiar situations in physical science<strong>and</strong> engineering. These examples have generally been fully worked, since, inthe authors’ experience, partially worked examples are unpopular with students.Only in a few cases, where trivial algebraic manipulation is involved, or whererepetition of the main text would result, has an example been left as an exercise<strong>for</strong> the reader. Nevertheless, a number of exercises also appear at the end of eachchapter, <strong>and</strong> these should give the reader ample opportunity to test his or herunderst<strong>and</strong>ing. Hints <strong>and</strong> answers to these exercises are also provided.With regard to the presentation of the mathematics, it has to be accepted thatmany equations (especially partial differential equations) can be written morecompactly by using subscripts, e.g. u xy <strong>for</strong> a second partial derivative, instead ofthe more familiar ∂ 2 u/∂x∂y, <strong>and</strong> that this certainly saves typographical space.However, <strong>for</strong> many students, the labour of mentally unpacking such equationsis sufficiently great that it is not possible to think of an equation’s physicalinterpretation at the same time. Consequently, wherever possible we have decidedto write out such expressions in their more obvious but longer <strong>for</strong>m.During the writing of this book we have received much help <strong>and</strong> encouragementfrom various colleagues at the Cavendish Laboratory, Clare College, Trinity Hall<strong>and</strong> Peterhouse. In particular, we would like to thank Peter Scheuer, whosecomments <strong>and</strong> general enthusiasm proved invaluable in the early stages. Forreading sections of the manuscript, <strong>for</strong> pointing out misprints <strong>and</strong> <strong>for</strong> numeroususeful comments, we thank many of our students <strong>and</strong> colleagues at the Universityof Cambridge. We are especially grateful to Chris Doran, John Huber, GarthLeder, Tom Körner <strong>and</strong>, not least, Mike Stobbs, who, sadly, died be<strong>for</strong>e the bookwas completed. We also extend our thanks to the University of Cambridge <strong>and</strong>the Cavendish teaching staff, whose examination questions <strong>and</strong> lecture h<strong>and</strong>-outshave collectively provided the basis <strong>for</strong> some of the examples included. Of course,any errors <strong>and</strong> ambiguities remaining are entirely the responsibility of the authors,<strong>and</strong> we would be most grateful to have them brought to our attention.We are indebted to Dave Green <strong>for</strong> a great deal of advice concerning typesettingin L A TEX <strong>and</strong> to Andrew Lovatt <strong>for</strong> various other computing tips. Our thanksalso go to Anja Visser <strong>and</strong> Graça Rocha <strong>for</strong> enduring many hours of (sometimesheated) debate. At Cambridge University Press, we are very grateful to our editorAdam Black <strong>for</strong> his help <strong>and</strong> patience <strong>and</strong> to Alison Woollatt <strong>for</strong> her experttypesetting of such a complicated text. We also thank our copy-editor SusanParkinson <strong>for</strong> many useful suggestions that have undoubtedly improved the styleof the book.Finally, on a personal note, KFR wishes to thank his wife Penny, not only <strong>for</strong>a long <strong>and</strong> happy marriage, but also <strong>for</strong> her support <strong>and</strong> underst<strong>and</strong>ing duringhis recent illness – <strong>and</strong> when things have not gone too well at the bridge table!MPH is indebted both to Rebecca Morris <strong>and</strong> to his parents <strong>for</strong> their tirelessxxvi


PREFACE TO THE FIRST EDITIONsupport <strong>and</strong> patience, <strong>and</strong> <strong>for</strong> their unending supplies of tea. SJB is grateful toAnthony Gritten <strong>for</strong> numerous relaxing discussions about J. S. Bach, to SusannahTicciati <strong>for</strong> her patience <strong>and</strong> underst<strong>and</strong>ing, <strong>and</strong> to Kate Isaak <strong>for</strong> her calminglate-night e-mails from the USA.Ken Riley, Michael Hobson <strong>and</strong> Stephen BenceCambridge, 1997xxvii


1Preliminary algebraThis opening chapter reviews the basic algebra of which a working knowledge ispresumed in the rest of the book. Many students will be familiar with much, ifnot all, of it, but recent changes in what is studied during secondary educationmean that it cannot be taken <strong>for</strong> granted that they will already have a masteryof all the topics presented here. The reader may assess which areas need furtherstudy or revision by attempting the exercises at the end of the chapter. The mainareas covered are polynomial equations <strong>and</strong> the related topic of partial fractions,curve sketching, coordinate geometry, trigonometric identities <strong>and</strong> the notions ofproof by induction or contradiction.1.1 Simple functions <strong>and</strong> equationsIt is normal practice when starting the mathematical investigation of a physicalproblem to assign an algebraic symbol to the quantity whose value is sought, eithernumerically or as an explicit algebraic expression. For the sake of definiteness, inthis chapter we will use x to denote this quantity most of the time. Subsequentsteps in the analysis involve applying a combination of known laws, consistencyconditions <strong>and</strong> (possibly) given constraints to derive one or more equationssatisfied by x. These equations may take many <strong>for</strong>ms, ranging from a simplepolynomial equation to, say, a partial differential equation with several boundaryconditions. Some of the more complicated possibilities are treated in the laterchapters of this book, but <strong>for</strong> the present we will be concerned with techniques<strong>for</strong> the solution of relatively straight<strong>for</strong>ward algebraic equations.1.1.1 Polynomials <strong>and</strong> polynomial equationsFirstly we consider the simplest type of equation, a polynomial equation, inwhicha polynomial expression in x, denoted by f(x), is set equal to zero <strong>and</strong> thereby1


PRELIMINARY ALGEBRA<strong>for</strong>ms an equation which is satisfied by particular values of x, called the roots ofthe equation:f(x) =a n x n + a n−1 x n−1 + ···+ a 1 x + a 0 =0. (1.1)Here n is an integer > 0, called the degree of both the polynomial <strong>and</strong> theequation, <strong>and</strong> the known coefficients a 0 ,a 1 ,...,a n are real quantities with a n ≠0.Equations such as (1.1) arise frequently in physical problems, the coefficients a ibeing determined by the physical properties of the system under study. What isneeded is to find some or all of the roots of (1.1), i.e. the x-values, α k ,thatsatisfyf(α k )=0;herek is an index that, as we shall see later, can take up to n differentvalues, i.e. k =1, 2,...,n. The roots of the polynomial equation can equally wellbe described as the zeros of the polynomial. When they are real, they correspondto the points at which a graph of f(x) crosses the x-axis. Roots that are complex(see chapter 3) do not have such a graphical interpretation.For polynomial equations containing powers of x greater than x 4 generalmethods do not exist <strong>for</strong> obtaining explicit expressions <strong>for</strong> the roots α k . Even<strong>for</strong> n = 3 <strong>and</strong> n = 4 the prescriptions <strong>for</strong> obtaining the roots are sufficientlycomplicated that it is usually preferable to obtain exact or approximate valuesby other methods. Only <strong>for</strong> n = 1 <strong>and</strong> n = 2 can closed-<strong>for</strong>m solutions be given.These results will be well known to the reader, but they are given here <strong>for</strong> thesake of completeness. For n = 1, (1.1) reduces to the linear equationa 1 x + a 0 = 0; (1.2)the solution (root) is α 1 = −a 0 /a 1 .Forn = 2, (1.1) reduces to the quadraticequationa 2 x 2 + a 1 x + a 0 = 0; (1.3)the two roots α 1 <strong>and</strong> α 2 are given by√α 1,2 = −a 1 ± a 2 1 − 4a 2a 0. (1.4)2a 2When discussing specifically quadratic equations, as opposed to more generalpolynomial equations, it is usual to write the equation in one of the two notationsax 2 + bx + c =0, ax 2 +2bx + c =0, (1.5)with respective explicit pairs of solutionsα 1,2 = −b ± √ b 2 − 4ac2a, α 1,2 = −b ± √ b 2 − ac. (1.6)aOf course, these two notations are entirely equivalent <strong>and</strong> the only importantpoint is to associate each <strong>for</strong>m of answer with the corresponding <strong>for</strong>m of equation;most people keep to one <strong>for</strong>m, to avoid any possible confusion.2


1.1 SIMPLE FUNCTIONS AND EQUATIONSIf the value of the quantity appearing under the square root sign is positivethen both roots are real; if it is negative then the roots <strong>for</strong>m a complex conjugatepair, i.e. they are of the <strong>for</strong>m p ± iq with p <strong>and</strong> q real (see chapter 3); if it haszero value then the two roots are equal <strong>and</strong> special considerations usually arise.Thus linear <strong>and</strong> quadratic equations can be dealt with in a cut-<strong>and</strong>-dried way.We now turn to methods <strong>for</strong> obtaining partial in<strong>for</strong>mation about the roots ofhigher-degree polynomial equations. In some circumstances the knowledge thatan equation has a root lying in a certain range, or that it has no real roots at all,is all that is actually required. For example, in the design of electronic circuitsit is necessary to know whether the current in a proposed circuit will breakinto spontaneous oscillation. To test this, it is sufficient to establish whether acertain polynomial equation, whose coefficients are determined by the physicalparameters of the circuit, has a root with a positive real part (see chapter 3);complete determination of all the roots is not needed <strong>for</strong> this purpose. If thecomplete set of roots of a polynomial equation is required, it can usually beobtained to any desired accuracy by numerical methods such as those describedin chapter 27.There is no explicit step-by-step approach to finding the roots of a generalpolynomial equation such as (1.1). In most cases analytic methods yield onlyin<strong>for</strong>mation about the roots, rather than their exact values. To explain the relevanttechniques we will consider a particular example, ‘thinking aloud’ on paper <strong>and</strong>exp<strong>and</strong>ing on special points about methods <strong>and</strong> lines of reasoning. In moreroutine situations such comment would be absent <strong>and</strong> the whole process briefer<strong>and</strong> more tightly focussed.Example: the cubic caseLet us investigate the roots of the equationg(x) =4x 3 +3x 2 − 6x − 1 = 0 (1.7)or, in an alternative phrasing, investigate the zeros of g(x).Wenotefirstofallthat this is a cubic equation. It can be seen that <strong>for</strong> x large <strong>and</strong> positive g(x)will be large <strong>and</strong> positive <strong>and</strong>, equally, that <strong>for</strong> x large <strong>and</strong> negative g(x) willbe large <strong>and</strong> negative. There<strong>for</strong>e, intuitively (or, more <strong>for</strong>mally, by continuity)g(x) must cross the x-axis at least once <strong>and</strong> so g(x) = 0 must have at least onereal root. Furthermore, it can be shown that if f(x) isannth-degree polynomialthen the graph of f(x) must cross the x-axis an even or odd number of timesas x varies between −∞ <strong>and</strong> +∞, according to whether n itself is even or odd.Thus a polynomial of odd degree always has at least one real root, but one ofeven degree may have no real root. A small complication, discussed later in thissection, occurs when repeated roots arise.Having established that g(x) = 0 has at least one real root, we may ask how3


PRELIMINARY ALGEBRAmany real roots it could have. To answer this we need one of the fundamentaltheorems of algebra, mentioned above:An nth-degree polynomial equation has exactly n roots.It should be noted that this does not imply that there are n real roots (only thatthere are not more than n); some of the roots may be of the <strong>for</strong>m p + iq.To make the above theorem plausible <strong>and</strong> to see what is meant by repeatedroots, let us suppose that the nth-degree polynomial equation f(x) = 0, (1.1), hasr roots α 1 ,α 2 ,...,α r , considered distinct <strong>for</strong> the moment. That is, we suppose thatf(α k )=0<strong>for</strong>k =1, 2,...,r,sothatf(x) vanishes only when x is equal to one ofthe r values α k . But the same can be said <strong>for</strong> the functionF(x) =A(x − α 1 )(x − α 2 ) ···(x − α r ), (1.8)in which A is a non-zero constant; F(x) can clearly be multiplied out to <strong>for</strong>m apolynomial expression.We now call upon a second fundamental result in algebra: that if two polynomialfunctions f(x) <strong>and</strong>F(x) have equal values <strong>for</strong> all values of x, then theircoefficients are equal on a term-by-term basis. In other words, we can equatethe coefficients of each <strong>and</strong> every power of x in the two expressions (1.8) <strong>and</strong>(1.1); in particular we can equate the coefficients of the highest power of x. Fromthis we have Ax r ≡ a n x n <strong>and</strong> thus that r = n <strong>and</strong> A = a n .Asr is both equalto n <strong>and</strong> to the number of roots of f(x) = 0, we conclude that the nth-degreepolynomial f(x) =0hasn roots. (Although this line of reasoning may make thetheorem plausible, it does not constitute a proof since we have not shown that itis permissible to write f(x) in the <strong>for</strong>m of equation (1.8).)We next note that the condition f(α k )=0<strong>for</strong>k =1, 2,...,r, could also be metif (1.8) were replaced byF(x) =A(x − α 1 ) m1 (x − α 2 ) m2 ···(x − α r ) mr , (1.9)with A = a n . In (1.9) the m k are integers ≥ 1 <strong>and</strong> are known as the multiplicitiesof the roots, m k being the multiplicity of α k . Exp<strong>and</strong>ing the right-h<strong>and</strong> side (RHS)leads to a polynomial of degree m 1 + m 2 + ···+ m r . This sum must be equal to n.Thus, if any of the m k is greater than unity then the number of distinct roots, r,is less than n; the total number of roots remains at n, but one or more of the α kcounts more than once. For example, the equationF(x) =A(x − α 1 ) 2 (x − α 2 ) 3 (x − α 3 )(x − α 4 )=0has exactly seven roots, α 1 being a double root <strong>and</strong> α 2 a triple root, whilst α 3 <strong>and</strong>α 4 are unrepeated (simple) roots.We can now say that our particular equation (1.7) has either one or three realroots but in the latter case it may be that not all the roots are distinct. To decidehow many real roots the equation has, we need to anticipate two ideas from the4


1.1 SIMPLE FUNCTIONS AND EQUATIONSφ 1 (x)φ 2 (x)β 2xβ 1 β 1β 2xFigure 1.1 Two curves φ 1 (x) <strong>and</strong>φ 2 (x), both with zero derivatives at thesame values of x, but with different numbers of real solutions to φ i (x) =0.next chapter. The first of these is the notion of the derivative of a function, <strong>and</strong>the second is a result known as Rolle’s theorem.The derivative f ′ (x) of a function f(x) measures the slope of the tangent tothe graph of f(x) at that value of x (see figure 2.1 in the next chapter). Forthe moment, the reader with no prior knowledge of calculus is asked to acceptthat the derivative of ax n is nax n−1 , so that the derivative g ′ (x) ofthecurveg(x) =4x 3 +3x 2 − 6x − 1 is given by g ′ (x) =12x 2 +6x − 6. Similar expressions<strong>for</strong> the derivatives of other polynomials are used later in this chapter.Rolle’s theorem states that if f(x) has equal values at two different values of xthen at some point between these two x-values its derivative is equal to zero; i.e.the tangent to its graph is parallel to the x-axis at that point (see figure 2.2).Having briefly mentioned the derivative of a function <strong>and</strong> Rolle’s theorem, wenow use them to establish whether g(x) has one or three real zeros. If g(x) =0does have three real roots α k ,i.e.g(α k )=0<strong>for</strong>k =1, 2, 3, then it follows fromRolle’s theorem that between any consecutive pair of them (say α 1 <strong>and</strong> α 2 )theremust be some real value of x at which g ′ (x) = 0. Similarly, there must be a furtherzero of g ′ (x) lying between α 2 <strong>and</strong> α 3 . Thus a necessary condition <strong>for</strong> three realroots of g(x) = 0 is that g ′ (x) = 0 itself has two real roots.However, this condition on the number of roots of g ′ (x) = 0, whilst necessary,is not sufficient to guarantee three real roots of g(x) = 0. This can be seen byinspecting the cubic curves in figure 1.1. For each of the two functions φ 1 (x) <strong>and</strong>φ 2 (x), the derivative is equal to zero at both x = β 1 <strong>and</strong> x = β 2 . Clearly, though,φ 2 (x) = 0 has three real roots whilst φ 1 (x) = 0 has only one. It is easy to see thatthe crucial difference is that φ 1 (β 1 )<strong>and</strong>φ 1 (β 2 ) have the same sign, whilst φ 2 (β 1 )<strong>and</strong> φ 2 (β 2 ) have opposite signs.It will be apparent that <strong>for</strong> some equations, φ(x) =0say,φ ′ (x) equals zero5


PRELIMINARY ALGEBRAat a value of x <strong>for</strong> which φ(x) is also zero. Then the graph of φ(x) justtouchesthe x-axis. When this happens the value of x so found is, in fact, a double realroot of the polynomial equation (corresponding to one of the m k in (1.9) havingthe value 2) <strong>and</strong> must be counted twice when determining the number of realroots.Finally, then, we are in a position to decide the number of real roots of theequationg(x) =4x 3 +3x 2 − 6x − 1=0.The equation g ′ (x) =0,withg ′ (x) =12x 2 +6x − 6, is a quadratic equation withexplicit solutions § β 1,2 = −3 ± √ 9+7212,so that β 1 = −1 <strong>and</strong>β 2 = 1 2 . The corresponding values of g(x) areg(β 1)=4<strong>and</strong>g(β 2 )=− 114 , which are of opposite sign. This indicates that 4x3 +3x 2 −6x−1 =0has three real roots, one lying in the range −1


1.1 SIMPLE FUNCTIONS AND EQUATIONS(iii) The equation f ′ (x) = 0 can be written as x(7x 5 +30x 4 +4x 2 − 3x +2)=0<strong>and</strong> thus x = 0 is a root. The derivative of f ′ (x), denoted by f ′′ (x), equals42x 5 + 150x 4 +12x 2 − 6x +2. That f ′ (x) is zero whilst f ′′ (x) is positiveat x = 0 indicates (subsection 2.1.8) that f(x) has a minimum there. This,together with the facts that f(0) is negative <strong>and</strong> f(∞) =∞, implies thatthe total number of real roots to the right of x = 0 must be odd. Sincethe total number of real roots must be odd, the number to the left mustbe even (0, 2, 4 or 6).This is about all that can be deduced by simple analytic methods in this case,although some further progress can be made in the ways indicated in exercise 1.3.There are, in fact, more sophisticated tests that examine the relative signs ofsuccessive terms in an equation such as (1.1), <strong>and</strong> in quantities derived fromthem, to place limits on the numbers <strong>and</strong> positions of roots. But they are notprerequisites <strong>for</strong> the remainder of this book <strong>and</strong> will not be pursued furtherhere.We conclude this section with a worked example which demonstrates that thepractical application of the ideas developed so far can be both short <strong>and</strong> decisive.◮For what values of k, ifany,doesf(x) =x 3 − 3x 2 +6x + k =0have three real roots?Firstly we study the equation f ′ (x) =0,i.e.3x 2 − 6x + 6 = 0. This is a quadratic equationbut, using (1.6), because 6 2 < 4 × 3 × 6, it can have no real roots. There<strong>for</strong>e, it followsimmediately that f(x) has no maximum or minimum; consequently f(x) = 0 cannot havemore than one real root, whatever the value of k. ◭1.1.2 Factorising polynomialsIn the previous subsection we saw how a polynomial with r given distinct zerosα k could be constructed as the product of factors containing those zeros:f(x) =a n (x − α 1 ) m1 (x − α 2 ) m2 ···(x − α r ) mr= a n x n + a n−1 x n−1 + ···+ a 1 x + a 0 , (1.10)with m 1 + m 2 + ···+ m r = n, the degree of the polynomial. It will cause no loss ofgenerality in what follows to suppose that all the zeros are simple, i.e. all m k =1<strong>and</strong> r = n, <strong>and</strong> this we will do.Sometimes it is desirable to be able to reverse this process, in particular whenone exact zero has been found by some method <strong>and</strong> the remaining zeros are tobe investigated. Suppose that we have located one zero, α; it is then possible towrite (1.10) asf(x) =(x − α)f 1 (x), (1.11)7


PRELIMINARY ALGEBRAwhere f 1 (x) is a polynomial of degree n−1. How can we find f 1 (x)? The procedureis much more complicated to describe in a general <strong>for</strong>m than to carry out <strong>for</strong>an equation with given numerical coefficients a i . If such manipulations are toocomplicated to be carried out mentally, they could be laid out along the lines ofan algebraic ‘long division’ sum. However, a more compact <strong>for</strong>m of calculationis as follows. Write f 1 (x) asf 1 (x) =b n−1 x n−1 + b n−2 x n−2 + b n−3 x n−3 + ···+ b 1 x + b 0 .Substitution of this <strong>for</strong>m into (1.11) <strong>and</strong> subsequent comparison of the coefficientsof x p <strong>for</strong> p = n, n − 1, ..., 1, 0 with those in the second line of (1.10) generatesthe series of equationsb n−1 = a n ,b n−2 − αb n−1 = a n−1 ,b n−3 − αb n−2 = a n−2 ,..b 0 − αb 1 = a 1 ,−αb 0 = a 0 .These can be solved successively <strong>for</strong> the b j , starting either from the top or fromthe bottom of the series. In either case the final equation used serves as a check;if it is not satisfied, at least one mistake has been made in the computation –or α is not a zero of f(x) = 0. We now illustrate this procedure with a workedexample.◮Determine by inspection the simple roots of the equationf(x) =3x 4 − x 3 − 10x 2 − 2x +4=0<strong>and</strong> hence, by factorisation, find the rest of its roots.From the pattern of coefficients it can be seen that x = −1 is a solution to the equation.We there<strong>for</strong>e writef(x) =(x +1)(b 3 x 3 + b 2 x 2 + b 1 x + b 0 ),whereb 3 =3,b 2 + b 3 = −1,b 1 + b 2 = −10,b 0 + b 1 = −2,b 0 =4.These equations give b 3 =3,b 2 = −4,b 1 = −6,b 0 = 4 (check) <strong>and</strong> sof(x) =(x +1)f 1 (x) =(x + 1)(3x 3 − 4x 2 − 6x +4).8


1.1 SIMPLE FUNCTIONS AND EQUATIONSWe now note that f 1 (x) =0ifx is set equal to 2. Thus x − 2isafactoroff 1 (x), whichthere<strong>for</strong>e can be written asf 1 (x) =(x − 2)f 2 (x) =(x − 2)(c 2 x 2 + c 1 x + c 0 )withc 2 =3,c 1 − 2c 2 = −4,c 0 − 2c 1 = −6,−2c 0 =4.These equations determine f 2 (x) as3x 2 +2x − 2. Since f 2 (x) = 0 is a quadratic equation,its solutions can be written explicitly asx = −1 ± √ 1+6.3Thus the four roots of f(x) =0are−1, 2, 1 (−1+√ 7) <strong>and</strong> 1 (−1 − √ 7). ◭3 31.1.3 Properties of rootsFrom the fact that a polynomial equation can be written in any of the alternative<strong>for</strong>msf(x) =a n x n + a n−1 x n−1 + ···+ a 1 x + a 0 =0,f(x) =a n (x − α 1 ) m1 (x − α 2 ) m2 ···(x − α r ) mr =0,f(x) =a n (x − α 1 )(x − α 2 ) ···(x − α n )=0,it follows that it must be possible to express the coefficients a i in terms of theroots α k . To take the most obvious example, comparison of the constant terms(<strong>for</strong>mally the coefficient of x 0 ) in the first <strong>and</strong> third expressions shows thata n (−α 1 )(−α 2 ) ···(−α n )=a 0 ,or, using the product notation,n∏α k =(−1) n a 0. (1.12)a nk=1Only slightly less obvious is a result obtained by comparing the coefficients ofx n−1 in the same two expressions of the polynomial:n∑α k = − a n−1. (1.13)a nk=1Comparing the coefficients of other powers of x yields further results, thoughthey are of less general use than the two just given. One such, which the readermay wish to derive, isn∑n∑j=1 k>jα j α k = a n−2a n. (1.14)9


PRELIMINARY ALGEBRAIn the case of a quadratic equation these root properties are used sufficientlyoften that they are worth stating explicitly, as follows. If the roots of the quadraticequation ax 2 + bx + c = 0 are α 1 <strong>and</strong> α 2 thenα 1 + α 2 = − b a ,α 1 α 2 = c a .If the alternative st<strong>and</strong>ard <strong>for</strong>m <strong>for</strong> the quadratic is used, b is replaced by 2b inboth the equation <strong>and</strong> the first of these results.◮Find a cubic equation whose roots are −4, 3 <strong>and</strong> 5.From results (1.12) – (1.14) we can compute that, arbitrarily setting a 3 =1,3∑3∑ 3∑3∏−a 2 = α k =4, a 1 = α j α k = −17, a 0 =(−1) 3 α k =60.k=1j=1 k>jThus a possible cubic equation is x 3 +(−4)x 2 +(−17)x + (60) = 0. Of course, any multipleof x 3 − 4x 2 − 17x + 60 = 0 will do just as well. ◭k=11.2 Trigonometric identitiesSo many of the applications of mathematics to physics <strong>and</strong> engineering areconcerned with periodic, <strong>and</strong> in particular sinusoidal, behaviour that a sure <strong>and</strong>ready h<strong>and</strong>ling of the corresponding mathematical functions is an essential skill.Even situations with no obvious periodicity are often expressed in terms ofperiodic functions <strong>for</strong> the purposes of analysis. Later in this book whole chaptersare devoted to developing the techniques involved, but as a necessary prerequisitewe here establish (or remind the reader of) some st<strong>and</strong>ard identities with which heor she should be fully familiar, so that the manipulation of expressions containingsinusoids becomes automatic <strong>and</strong> reliable. So as to emphasise the angular natureof the argument of a sinusoid we will denote it in this section by θ rather than x.1.2.1 Single-angle identitiesWe give without proof the basic identity satisfied by the sinusoidal functions sin θ<strong>and</strong> cos θ, namelycos 2 θ +sin 2 θ =1. (1.15)If sin θ <strong>and</strong> cos θ have been defined geometrically in terms of the coordinates ofa point on a circle, a reference to the name of Pythagoras will suffice to establishthis result. If they have been defined by means of series (with θ expressed inradians) then the reader should refer to Euler’s equation (3.23) on page 93, <strong>and</strong>note that e iθ has unit modulus if θ is real.10


1.2 TRIGONOMETRIC IDENTITIESy ′yRPMTNBx ′OAxFigure 1.2 Illustration of the compound-angle identities. Refer to the maintext <strong>for</strong> details.Other st<strong>and</strong>ard single-angle <strong>for</strong>mulae derived from (1.15) by dividing throughby various powers of sin θ <strong>and</strong> cos θ are1+tan 2 θ =sec 2 θ, (1.16)cot 2 θ +1=cosec 2 θ. (1.17)1.2.2 Compound-angle identitiesThe basis <strong>for</strong> building expressions <strong>for</strong> the sinusoidal functions of compoundangles are those <strong>for</strong> the sum <strong>and</strong> difference of just two angles, since all othercases can be built up from these, in principle. Later we will see that a study ofcomplex numbers can provide a more efficient approach in some cases.To prove the basic <strong>for</strong>mulae <strong>for</strong> the sine <strong>and</strong> cosine of a compound angleA + B in terms of the sines <strong>and</strong> cosines of A <strong>and</strong> B, we consider the constructionshown in figure 1.2. It shows two sets of axes, Oxy <strong>and</strong> Ox ′ y ′ , with a commonorigin but rotated with respect to each other through an angle A. The pointP lies on the unit circle centred on the common origin O <strong>and</strong> has coordinatescos(A + B), sin(A + B) with respect to the axes Oxy <strong>and</strong> coordinates cos B,sin Bwith respect to the axes Ox ′ y ′ .Parallels to the axes Oxy (dotted lines) <strong>and</strong> Ox ′ y ′ (broken lines) have beendrawn through P . Further parallels (MR <strong>and</strong> RN) totheOx ′ y ′ axes have been11


PRELIMINARY ALGEBRAdrawn through R, the point (0, sin(A + B)) in the Oxy system. That all the anglesmarked with the symbol • are equal to A follows from the simple geometry ofright-angled triangles <strong>and</strong> crossing lines.We now determine the coordinates of P in terms of lengths in the figure,expressing those lengths in terms of both sets of coordinates:(i) cos B = x ′ = TN + NP = MR + NP= OR sin A + RP cos A = sin(A + B)sinA +cos(A + B)cosA;(ii) sin B = y ′ = OM − TM = OM − NR= OR cos A − RP sin A = sin(A + B)cosA − cos(A + B)sinA.Now, if equation (i) is multiplied by sin A <strong>and</strong> added to equation (ii) multipliedby cos A, the result issin A cos B +cosA sin B = sin(A + B)(sin 2 A +cos 2 A)=sin(A + B).Similarly, if equation (ii) is multiplied by sin A <strong>and</strong> subtracted from equation (i)multiplied by cos A, the result iscos A cos B − sin A sin B =cos(A + B)(cos 2 A +sin 2 A)=cos(A + B).Corresponding graphically based results can be derived <strong>for</strong> the sines <strong>and</strong> cosinesof the difference of two angles; however, they are more easily obtained by settingB to −B in the previous results <strong>and</strong> remembering that sin B becomes − sin Bwhilst cos B is unchanged. The four results may be summarised bysin(A ± B) =sinA cos B ± cos A sin B (1.18)cos(A ± B) =cosA cos B ∓ sin A sin B. (1.19)St<strong>and</strong>ard results can be deduced from these by setting one of the two anglesequal to π or to π/2:sin(π − θ) =sinθ, cos(π − θ) =− cos θ, (1.20)sin ( 12 π − θ) =cosθ, cos ( 12 π − θ) =sinθ, (1.21)From these basic results many more can be derived. An immediate deduction,obtained by taking the ratio of the two equations (1.18) <strong>and</strong> (1.19) <strong>and</strong> thendividing both the numerator <strong>and</strong> denominator of this ratio by cos A cos B, istan A ± tan Btan(A ± B) =1 ∓ tan A tan B . (1.22)One application of this result is a test <strong>for</strong> whether two lines on a graphare orthogonal (perpendicular); more generally, it determines the angle betweenthem. The st<strong>and</strong>ard notation <strong>for</strong> a straight-line graph is y = mx + c, inwhichmis the slope of the graph <strong>and</strong> c is its intercept on the y-axis. It should be notedthat the slope m is also the tangent of the angle the line makes with the x-axis.12


1.2 TRIGONOMETRIC IDENTITIESConsequently the angle θ 12 between two such straight-line graphs is equal to thedifference in the angles they individually make with the x-axis, <strong>and</strong> the tangentof that angle is given by (1.22):tan θ 12 = tan θ 1 − tan θ 21+tanθ 1 tan θ 2= m 1 − m 21+m 1 m 2. (1.23)For the lines to be orthogonal we must have θ 12 = π/2, i.e. the final fraction onthe RHS of the above equation must equal ∞, <strong>and</strong>som 1 m 2 = −1. (1.24)A kind of inversion of equations (1.18) <strong>and</strong> (1.19) enables the sum or differenceof two sines or cosines to be expressed as the product of two sinusoids; theprocedure is typified by the following. Adding together the expressions given by(1.18) <strong>for</strong> sin(A + B) <strong>and</strong> sin(A − B) yieldssin(A + B)+sin(A − B) =2sinA cos B.If we now write A + B = C <strong>and</strong> A − B = D, this becomes( ) C + Dsin C +sinD =2sin cos2( C − DIn a similar way each of the following equations can be derived:2). (1.25)( ) ( )C + D C − Dsin C − sin D =2cos sin , (1.26)22( ) ( )C + D C − Dcos C +cosD =2cos cos , (1.27)22( ) ( )C + D C − Dcos C − cos D = −2sin sin . (1.28)22The minus sign on the right of the last of these equations should be noted; it mayhelp to avoid overlooking this ‘oddity’ to recall that if C>Dthen cos C


PRELIMINARY ALGEBRA<strong>and</strong> use made of equation (1.15), the following results are obtained:sin 2θ =2sinθ cos θ, (1.29)cos 2θ =cos 2 θ − sin 2 θ=2cos 2 θ − 1=1− 2sin 2 θ, (1.30)tan 2θ =2tanθ1 − tan 2 θ . (1.31)A further set of identities enables sinusoidal functions of θ to be expressed interms of polynomial functions of a variable t = tan(θ/2). They are not used intheir primary role until the next chapter, but we give a derivation of them here<strong>for</strong> reference.If t = tan(θ/2), then it follows from (1.16) that 1+t 2 =sec 2 (θ/2) <strong>and</strong> cos(θ/2) =(1 + t 2 ) −1/2 , whilst sin(θ/2) = t(1 + t 2 ) −1/2 . Now, using (1.29) <strong>and</strong> (1.30), we maywrite:sin θ =2sin θ 2 cos θ 2 = 2t1+t 2 , (1.32)cos θ =cos 2 θ 2 − θ sin2 2 = 1 − t21+t 2 , (1.33)tan θ =2t1 − t 2 . (1.34)It can be further shown that the derivative of θ with respect to t takes thealgebraic <strong>for</strong>m 2/(1 + t 2 ). This completes a package of results that enablesexpressions involving sinusoids, particularly when they appear as integr<strong>and</strong>s, tobe cast in more convenient algebraic <strong>for</strong>ms. The proof of the derivative property<strong>and</strong> examples of use of the above results are given in subsection (2.2.7).We conclude this section with a worked example which is of such a commonlyoccurring <strong>for</strong>m that it might be considered a st<strong>and</strong>ard procedure.◮Solve <strong>for</strong> θ the equationa sin θ + b cos θ = k,where a, b <strong>and</strong> k are given real quantities.To solve this equation we make use of result (1.18) by setting a = K cos φ <strong>and</strong> b = K sin φ<strong>for</strong> suitable values of K <strong>and</strong> φ. We then havek = K cos φ sin θ + K sin φ cos θ = K sin(θ + φ),withK 2 = a 2 + b 2 <strong>and</strong> φ =tan −1 b a .Whether φ lies in 0 ≤ φ ≤ π or in −π


1.3 COORDINATE GEOMETRYwith K <strong>and</strong> φ as given above. Notice that the inverse sine yields two values in the range 0to 2π <strong>and</strong> that there is no real solution to the original equation if |k| > |K| =(a 2 +b 2 ) 1/2 . ◭1.3 Coordinate geometryWe have already mentioned the st<strong>and</strong>ard <strong>for</strong>m <strong>for</strong> a straight-line graph, namelyy = mx + c, (1.35)representing a linear relationship between the independent variable x <strong>and</strong> thedependent variable y. Theslopem is equal to the tangent of the angle the linemakes with the x-axis whilst c is the intercept on the y-axis.An alternative <strong>for</strong>m <strong>for</strong> the equation of a straight line isto which (1.35) is clearly connected byax + by + k =0, (1.36)m = − a b<strong>and</strong> c = − k b .This <strong>for</strong>m treats x <strong>and</strong> y on a more symmetrical basis, the intercepts on the twoaxes being −k/a <strong>and</strong> −k/b respectively.A power relationship between two variables, i.e. one of the <strong>for</strong>m y = Ax n ,canalso be cast into straight-line <strong>for</strong>m by taking the logarithms of both sides. Whilstit is normal in mathematical work to use natural logarithms (to base e, writtenln x), <strong>for</strong> practical investigations logarithms to base 10 are often employed. Ineither case the <strong>for</strong>m is the same, but it needs to be remembered which has beenused when recovering the value of A from fitted data. In the mathematical (basee) <strong>for</strong>m, the power relationship becomesln y = n ln x +lnA. (1.37)Now the slope gives the power n, whilst the intercept on the ln y axis is ln A,which yields A, either by exponentiation or by taking antilogarithms.The other st<strong>and</strong>ard coordinate <strong>for</strong>ms of two-dimensional curves that studentsshould know <strong>and</strong> recognise are those concerned with the conic sections – so calledbecause they can all be obtained by taking suitable sections across a (double)cone. Because the conic sections can take many different orientations <strong>and</strong> scalingstheir general <strong>for</strong>m is complex,Ax 2 + By 2 + Cxy + Dx + Ey + F =0, (1.38)but each can be represented by one of four generic <strong>for</strong>ms, an ellipse, a parabola, ahyperbola or, the degenerate <strong>for</strong>m, a pair of straight lines. If they are reduced totheir st<strong>and</strong>ard representations, in which axes of symmetry are made to coincide15


PRELIMINARY ALGEBRAwith the coordinate axes, the first three take the <strong>for</strong>ms(x − α) 2 (y − β)2a 2 +b 2 = 1 (ellipse), (1.39)(y − β) 2 =4a(x − α) (parabola), (1.40)(x − α) 2 (y − β)2a 2 −b 2 = 1 (hyperbola). (1.41)Here, (α, β) gives the position of the ‘centre’ of the curve, usually taken asthe origin (0, 0) when this does not conflict with any imposed conditions. Theparabola equation given is that <strong>for</strong> a curve symmetric about a line parallel tothe x-axis. For one symmetrical about a parallel to the y-axis the equation wouldread (x − α) 2 =4a(y − β).Of course, the circle is the special case of an ellipse in which b = a <strong>and</strong> theequation takes the <strong>for</strong>m(x − α) 2 +(y − β) 2 = a 2 . (1.42)The distinguishing characteristic of this equation is that when it is expressed inthe <strong>for</strong>m (1.38) the coefficients of x 2 <strong>and</strong> y 2 are equal <strong>and</strong> that of xy is zero; thisproperty is not changed by any reorientation or scaling <strong>and</strong> so acts to identify ageneral conic as a circle.Definitions of the conic sections in terms of geometrical properties are alsoavailable; <strong>for</strong> example, a parabola can be defined as the locus of a point thatis always at the same distance from a given straight line (the directrix) asitisfrom a given point (the focus). When these properties are expressed in Cartesiancoordinates the above equations are obtained. For a circle, the defining propertyis that all points on the curve are a distance a from (α, β); (1.42) expresses thisrequirement very directly. In the following worked example we derive the equation<strong>for</strong> a parabola.◮Find the equation of a parabola that has the line x = −a as its directrix <strong>and</strong> the point(a, 0) as its focus.Figure 1.3 shows the situation in Cartesian coordinates. Expressing the defining requirementthat PN <strong>and</strong> PF are equal in length gives(x + a) =[(x − a) 2 + y 2 ] 1/2 ⇒ (x + a) 2 =(x − a) 2 + y 2which, on expansion of the squared terms, immediately gives y 2 =4ax. This is (1.40) withα <strong>and</strong> β both set equal to zero. ◭Although the algebra is more complicated, the same method can be used toderive the equations <strong>for</strong> the ellipse <strong>and</strong> the hyperbola. In these cases the distancefrom the fixed point is a definite fraction, e, known as the eccentricity, ofthedistance from the fixed line. For an ellipse 0


1.3 COORDINATE GEOMETRYyNP(x, y)OF(a, 0)xx = −aFigure 1.3 Construction of a parabola using the point (a, 0) as the focus <strong>and</strong>the line x = −a as the directrix.The values of a <strong>and</strong> b (with a ≥ b) in equation (1.39) <strong>for</strong> an ellipse are relatedto e throughe 2 = a2 − b 2a 2<strong>and</strong> give the lengths of the semi-axes of the ellipse. If the ellipse is centred onthe origin, i.e. α = β = 0, then the focus is (−ae, 0) <strong>and</strong> the directrix is the linex = −a/e.For each conic section curve, although we have two variables, x <strong>and</strong> y, theyarenot independent, since if one is given then the other can be determined. However,determining y when x is given, say, involves solving a quadratic equation on eachoccasion, <strong>and</strong> so it is convenient to have parametric representations of the curves.A parametric representation allows each point on a curve to be associated witha unique value of a single parameter t. The simplest parametric representations<strong>for</strong> the conic sections are as given below, though that <strong>for</strong> the hyperbola useshyperbolic functions, not <strong>for</strong>mally introduced until chapter 3. That they do givevalid parameterizations can be verified by substituting them into the st<strong>and</strong>ard<strong>for</strong>ms (1.39)–(1.41); in each case the st<strong>and</strong>ard <strong>for</strong>m is reduced to an algebraic ortrigonometric identity.x = α + a cos φ, y = β + b sin φ (ellipse),x = α + at 2 , y = β +2at (parabola),x = α + a cosh φ, y = β + b sinh φ (hyperbola).As a final example illustrating several topics from this section we now prove17


PRELIMINARY ALGEBRAthe well-known result that the angle subtended by a diameter at any point on acircle is a right angle.◮Taking the diameter to be the line joining Q =(−a, 0) <strong>and</strong> R =(a, 0) <strong>and</strong> the point P tobe any point on the circle x 2 + y 2 = a 2 , prove that angle QP R is a right angle.If P is the point (x, y), the slope of the line QP ism 1 = y − 0x − (−a) = yx + a .That of RP ism 2 = y − 0x − (a) = yx − a .Thusm 1 m 2 =y2x 2 − a . 2But, since P is on the circle, y 2 = a 2 − x 2 <strong>and</strong> consequently m 1 m 2 = −1. From result (1.24)this implies that QP <strong>and</strong> RP are orthogonal <strong>and</strong> that QP R is there<strong>for</strong>e a right angle. Notethat this is true <strong>for</strong> any point P on the circle. ◭1.4 Partial fractionsIn subsequent chapters, <strong>and</strong> in particular when we come to study integrationin chapter 2, we will need to express a function f(x) that is the ratio of twopolynomials in a more manageable <strong>for</strong>m. To remove some potential complexityfrom our discussion we will assume that all the coefficients in the polynomialsare real, although this is not an essential simplification.The behaviour of f(x) is crucially determined by the location of the zeros ofits denominator, i.e. if f(x) is written as f(x) =g(x)/h(x) where both g(x) <strong>and</strong>h(x) are polynomials, § then f(x) changes extremely rapidly when x is close tothose values α i that are the roots of h(x) = 0. To make such behaviour explicit,we write f(x) as a sum of terms such as A/(x − α) n ,inwhichA is a constant, α isone of the α i that satisfy h(α i )=0<strong>and</strong>n is a positive integer. Writing a functionin this way is known as expressing it in partial fractions.Suppose, <strong>for</strong> the sake of definiteness, that we wish to express the functionf(x) =4x +2x 2 +3x +2§ It is assumed that the ratio has been reduced so that g(x) <strong>and</strong>h(x) do not contain any commonfactors, i.e. there is no value of x that makes both vanish at the same time. We may also assumewithout any loss of generality that the coefficient of the highest power of x in h(x) has been madeequal to unity, if necessary, by dividing both numerator <strong>and</strong> denominator by the coefficient of thishighest power.18


1.4 PARTIAL FRACTIONSin partial fractions, i.e. to write it asf(x) = g(x) 4x +2=h(x) x 2 +3x +2 = A 1(x − α 1 ) + A 2n1 (x − α 2 ) + ··· . n2 (1.43)The first question that arises is that of how many terms there should be onthe right-h<strong>and</strong> side (RHS). Although some complications occur when h(x) hasrepeated roots (these are considered below) it is clear that f(x) only becomesinfinite at the two values of x, α 1 <strong>and</strong> α 2 , that make h(x) = 0. Consequently theRHS can only become infinite at the same two values of x <strong>and</strong> there<strong>for</strong>e containsonly two partial fractions – these are the ones shown explicitly. This argumentcan be trivially extended (again temporarily ignoring the possibility of repeatedroots of h(x)) to show that if h(x) is a polynomial of degree n then there should ben terms on the RHS, each containing a different root α i of the equation h(α i )=0.A second general question concerns the appropriate values of the n i .Thisisanswered by putting the RHS over a common denominator, which will clearlyhave to be the product (x − α 1 ) n1 (x − α 2 ) n2 ···. Comparison of the highest powerof x in this new RHS with the same power in h(x) showsthatn 1 + n 2 + ···= n.This result holds whether or not h(x) = 0 has repeated roots <strong>and</strong>, although wedo not give a rigorous proof, strongly suggests the following correct conclusions.• The number of terms on the RHS is equal to the number of distinct roots ofh(x) = 0, each term having a different root α i in its denominator (x − α i ) ni .• If α i is a multiple root of h(x) = 0 then the value to be assigned to n i in (1.43) isthat of m i when h(x) is written in the product <strong>for</strong>m (1.9). Further, as discussedon p. 23, A i has to be replaced by a polynomial of degree m i − 1. This is also<strong>for</strong>mally true <strong>for</strong> non-repeated roots, since then both m i <strong>and</strong> n i are equal tounity.Returning to our specific example we note that the denominator h(x) has zerosat x = α 1 = −1 <strong>and</strong>x = α 2 = −2; these x-values are the simple (non-repeated)roots of h(x) = 0. Thus the partial fraction expansion will be of the <strong>for</strong>m4x +2x 2 +3x +2 = A 1x +1 + A 2x +2 . (1.44)We now list several methods available <strong>for</strong> determining the coefficients A 1 <strong>and</strong>A 2 . We also remind the reader that, as with all the explicit examples <strong>and</strong> techniquesdescribed, these methods are to be considered as models <strong>for</strong> the h<strong>and</strong>ling of anyratio of polynomials, with or without characteristics that make it a special case.(i) The RHS can be put over a common denominator, in this case (x+1)(x+2),<strong>and</strong> then the coefficients of the various powers of x can be equated in the19


PRELIMINARY ALGEBRAnumerators on both sides of the equation. This leads to4x +2=A 1 (x +2)+A 2 (x +1),4=A 1 + A 2 2=2A 1 + A 2 .Solving the simultaneous equations <strong>for</strong> A 1 <strong>and</strong> A 2 gives A 1 = −2 <strong>and</strong>A 2 =6.(ii) A second method is to substitute two (or more generally n) differentvalues of x into each side of (1.44) <strong>and</strong> so obtain two (or n) simultaneousequations <strong>for</strong> the two (or n) constantsA i . To justify this practical way ofproceeding it is necessary, strictly speaking, to appeal to method (i) above,which establishes that there are unique values <strong>for</strong> A 1 <strong>and</strong> A 2 valid <strong>for</strong>all values of x. It is normally very convenient to take zero as one of thevalues of x, but of course any set will do. Suppose in the present case thatwe use the values x = 0 <strong>and</strong> x = 1 <strong>and</strong> substitute in (1.44). The resultingequations are22 = A 11 + A 22 ,66 = A 12 + A 23 ,which on solution give A 1 = −2 <strong>and</strong>A 2 = 6, as be<strong>for</strong>e. The reader caneasily verify that any other pair of values <strong>for</strong> x (except <strong>for</strong> a pair thatincludes α 1 or α 2 ) gives the same values <strong>for</strong> A 1 <strong>and</strong> A 2 .(iii) The very reason why method (ii) fails if x is chosen as one of the rootsα i of h(x) = 0 can be made the basis <strong>for</strong> determining the values of the A icorresponding to non-multiple roots without having to solve simultaneousequations. The method is conceptually more difficult than the other methodspresented here, <strong>and</strong> needs results from the theory of complex variables(chapter 24) to justify it. However, we give a practical ‘cookbook’ recipe<strong>for</strong> determining the coefficients.(a) To determine the coefficient A k , imagine the denominator h(x)written as the product (x − α 1 )(x − α 2 ) ···(x − α n ), with any m-foldrepeated root giving rise to m factors in parentheses.(b) Now set x equal to α k <strong>and</strong> evaluate the expression obtained afteromitting the factor that reads α k − α k .(c) Divide the value so obtained into g(α k ); the result is the requiredcoefficient A k .For our specific example we find that in step (a) that h(x) =(x +1)(x +2)<strong>and</strong> that in evaluating A 1 step (b) yields −1 + 2, i.e. 1. Since g(−1) =4(−1) + 2 = −2, step (c) gives A 1 as (−2)/(1), i.e in agreement with ourother evaluations. In a similar way A 2 is evaluated as (−6)/(−1) = 6.20


1.4 PARTIAL FRACTIONSThus any one of the methods listed above shows that4x +2x 2 +3x +2 = −2x +1 + 6x +2 .The best method to use in any particular circumstance will depend on thecomplexity, in terms of the degrees of the polynomials <strong>and</strong> the multiplicities ofthe roots of the denominator, of the function being considered <strong>and</strong>, to someextent, on the individual inclinations of the student; some prefer lengthy butstraight<strong>for</strong>ward solution of simultaneous equations, whilst others feel more athome carrying through shorter but more abstract calculations in their heads.1.4.1 Complications <strong>and</strong> special casesHaving established the basic method <strong>for</strong> partial fractions, we now show, throughfurther worked examples, how some complications are dealt with by extensionsto the procedure. These extensions are introduced one at a time, but of course inany practical application more than one may be involved.The degree of the numerator is greater than or equal to that of the denominatorAlthough we have not specifically mentioned the fact, it will be apparent fromtrying to apply method (i) of the previous subsection to such a case, that if thedegree of the numerator (m) is not less than that of the denominator (n) then theratio of two polynomials cannot be expressed in partial fractions.To get round this difficulty it is necessary to start by dividing the denominatorh(x) into the numerator g(x) to obtain a further polynomial, which we will denoteby s(x), together with a function t(x) thatis a ratio of two polynomials <strong>for</strong> whichthe degree of the numerator is less than that of the denominator. The functiont(x) can there<strong>for</strong>e be exp<strong>and</strong>ed in partial fractions. As a <strong>for</strong>mula,f(x) = g(x)r(x)= s(x)+t(x) ≡ s(x)+h(x) h(x) . (1.45)It is apparent that the polynomial r(x) istheremainder obtained when g(x) isdivided by h(x), <strong>and</strong>, in general, will be a polynomial of degree n − 1. It is alsoclear that the polynomial s(x) will be of degree m − n. Again, the actual divisionprocess can be set out as an algebraic long division sum but is probably moreeasily h<strong>and</strong>led by writing (1.45) in the <strong>for</strong>mor, more explicitly, asg(x) =s(x)h(x)+r(x) (1.46)g(x) =(s m−n x m−n + s m−n−1 x m−n−1 + ···+ s 0 )h(x)+(r n−1 x n−1 + r n−2 x n−2 + ···+ r 0 )(1.47)<strong>and</strong> then equating coefficients.21


PRELIMINARY ALGEBRAWe illustrate this procedure with the following worked example.◮Find the partial fraction decomposition of the functionf(x) = x3 +3x 2 +2x +1.x 2 − x − 6Since the degree of the numerator is 3 <strong>and</strong> that of the denominator is 2, a preliminarylong division is necessary. The polynomial s(x) resulting from the division will have degree3 − 2 = 1 <strong>and</strong> the remainder r(x) will be of degree 2 − 1 = 1 (or less). Thus we writex 3 +3x 2 +2x +1=(s 1 x + s 0 )(x 2 − x − 6) + (r 1 x + r 0 ).From equating the coefficients of the various powers of x on the two sides of the equation,starting with the highest, we now obtain the simultaneous equations1=s 1 ,3=s 0 − s 1 ,2=−s 0 − 6s 1 + r 1 ,1=−6s 0 + r 0 .These are readily solved, in the given order, to yield s 1 =1,s 0 =4,r 1 =12<strong>and</strong>r 0 = 25.Thus f(x) can be written as12x +25f(x) =x +4+x 2 − x − 6 .The last term can now be decomposed into partial fractions as previously. The zeros ofthe denominator are at x =3<strong>and</strong>x = −2 <strong>and</strong> the application of any method from theprevious subsection yields the respective constants as A 1 =12 1 <strong>and</strong> A 5 2 = − 1 . Thus the5final partial fraction decomposition of f(x) isx +4+615(x − 3) − 15(x +2) . ◭Factors of the <strong>for</strong>m a 2 + x 2 in the denominatorWe have so far assumed that the roots of h(x) = 0, needed <strong>for</strong> the factorisation ofthe denominator of f(x), can always be found. In principle they always can butin some cases they are not real. Consider, <strong>for</strong> example, attempting to express inpartial fractions a polynomial ratio whose denominator is h(x) =x 3 −x 2 +2x−2.Clearly x = 1 is a zero of h(x), <strong>and</strong> so a first factorisation is (x − 1)(x 2 +2).However we cannot make any further progress because the factor x 2 + 2 cannotbe expressed as (x − α)(x − β) <strong>for</strong>anyrealα <strong>and</strong> β.Complex numbers are introduced later in this book (chapter 3) <strong>and</strong>, when thereader has studied them, he or she may wish to justify the procedure set outbelow. It can be shown to be equivalent to that already given, but the zeros ofh(x) are now allowed to be complex <strong>and</strong> terms that are complex conjugates ofeach other are combined to leave only real terms.Since quadratic factors of the <strong>for</strong>m a 2 +x 2 that appear in h(x) cannot be reducedto the product of two linear factors, partial fraction expansions including themneed to have numerators in the corresponding terms that are not simply constants22


1.4 PARTIAL FRACTIONSA i but linear functions of x, i.e.ofthe<strong>for</strong>mB i x + C i . Thus, in the expansion,linear terms (first-degree polynomials) in the denominator have constants (zerodegreepolynomials) in their numerators, whilst quadratic terms (second-degreepolynomials) in the denominator have linear terms (first-degree polynomials) intheir numerators. As a symbolic <strong>for</strong>mula, the partial fraction expansion ofg(x)(x − α 1 )(x − α 2 ) ···(x − α p )(x 2 + a 2 1 )(x2 + a 2 2 ) ···(x2 + a 2 q )should take the <strong>for</strong>mA 1x − α 1+ A 2x − α 2+ ···+A p+ B 1x + C 1x − α p x 2 + a 2 + B 2x + C 21x 2 + a 2 2+ ···+ B qx + C qx 2 + a 2 .qOf course, the degree of g(x) must be less than p +2q; if it is not, an initialdivision must be carried out as demonstrated earlier.Repeated factors in the denominatorConsider trying (incorrectly) to exp<strong>and</strong>f(x) =x − 4(x +1)(x − 2) 2in partial fraction <strong>for</strong>m as follows:x − 4(x +1)(x − 2) 2 = A 1x +1 + A 2(x − 2) 2 .Multiplying both sides of this supposed equality by (x +1)(x − 2) 2 produces anequation whose LHS is linear in x, whilst its RHS is quadratic. This is clearlywrong <strong>and</strong> so an expansion in the above <strong>for</strong>m cannot be valid. The correction wemust make is very similar to that needed in the previous subsection, namely thatsince (x − 2) 2 is a quadratic polynomial the numerator of the term containing itmust be a first-degree polynomial, <strong>and</strong> not simply a constant.The correct <strong>for</strong>m <strong>for</strong> the part of the expansion containing the doubly repeatedroot is there<strong>for</strong>e (Bx+ C)/(x − 2) 2 . Using this <strong>for</strong>m <strong>and</strong> either of methods (i) <strong>and</strong>(ii) <strong>for</strong> determining the constants gives the full partial fraction expansion asx − 4(x +1)(x − 2) 2 = − 5 5x − 16+9(x +1) 9(x − 2) 2 ,as the reader may verify.Since any term of the <strong>for</strong>m (Bx + C)/(x − α) 2 can be written asB(x − α)+C + Bα(x − α) 2 = Bx − α + C + Bα(x − α) 2 ,<strong>and</strong> similarly <strong>for</strong> multiply repeated roots, an alternative <strong>for</strong>m <strong>for</strong> the part of thepartial fraction expansion containing a repeated root α isD 1x − α + D 2(x − α) 2 + ···+ D p(x − α) p . (1.48)23


PRELIMINARY ALGEBRAIn this <strong>for</strong>m, all x-dependence has disappeared from the numerators but at theexpense of p−1 additional terms; the total number of constants to be determinedremains unchanged, as it must.When describing possible methods of determining the constants in a partialfraction expansion, we noted that method (iii), p. 20, which avoids the need tosolve simultaneous equations, is restricted to terms involving non-repeated roots.In fact, it can be applied in repeated-root situations, when the expansion is putin the <strong>for</strong>m (1.48), but only to find the constant in the term involving the largestinverse power of x − α, i.e.D p in (1.48).We conclude this section with a more protracted worked example that containsall three of the complications discussed.◮Resolve the following expression F(x) into partial fractions:F(x) = x5 − 2x 4 − x 3 +5x 2 − 46x + 100.(x 2 +6)(x − 2) 2We note that the degree of the denominator (4) is not greater than that of the numerator(5), <strong>and</strong> so we must start by dividing the latter by the <strong>for</strong>mer. It follows, from the differencein degrees <strong>and</strong> the coefficients of the highest powers in each, that the result will be a linearexpression s 1 x + s 0 with the coefficient s 1 equal to 1. Thus the numerator of F(x) mustbeexpressible as(x + s 0 )(x 4 − 4x 3 +10x 2 − 24x + 24) + (r 3 x 3 + r 2 x 2 + r 1 x + r 0 ),where the second factor in parentheses is the denominator of F(x) written as a polynomial.Equating the coefficients of x 4 gives −2 =−4+s 0 <strong>and</strong> fixes s 0 as 2. Equating the coefficientsof powers less than 4 gives equations involving the coefficients r i as follows:−1 =−8+10+r 3 ,5=−24+20+r 2 ,−46 = 24 − 48 + r 1 ,100 = 48 + r 0 .Thus the remainder polynomial r(x) can be constructed <strong>and</strong> F(x) written asF(x) =x +2+ −3x3 +9x 2 − 22x +52≡ x +2+f(x).(x 2 +6)(x − 2) 2The polynomial ratio f(x) can now be expressed in partial fraction <strong>for</strong>m, noting that itsdenominator contains both a term of the <strong>for</strong>m x 2 + a 2 <strong>and</strong> a repeated root. Thusf(x) = Bx + Cx 2 +6 + D 1x − 2 + D 2(x − 2) . 2We could now put the RHS of this equation over the common denominator (x 2 +6)(x−2) 2<strong>and</strong> find B,C,D 1 <strong>and</strong> D 2 by equating coefficients of powers of x. It is quicker, however,to use methods (iii) <strong>and</strong> (ii). Method (iii) gives D 2 as (−24 + 36 − 44 + 52)/(4 + 6) = 2.We choose to evaluate the other coefficients by method (ii), <strong>and</strong> setting x =0,x =1<strong>and</strong>24


1.5 BINOMIAL EXPANSIONx = −1 gives respectively5224 = C 6 − D 12 + 2 4 ,367 = B + C − D 1 +2,78663 = C − B − D 17 3 + 2 9 .These equations reduce to4C − 12D 1 =40,B + C − 7D 1 =22,−9B +9C − 21D 1 =72,with solution B =0,C =1,D 1 = −3.Thus, finally, we may rewrite the original expression F(x) in partial fractions asF(x) =x +2+ 1x 2 +6 − 3x − 2 + 2(x − 2) . ◭ 21.5 Binomial expansionEarlier in this chapter we were led to consider functions containing powers ofthe sum or difference of two terms, e.g. (x − α) m . Later in this book we will findnumerous occasions on which we wish to write such a product of repeated factorsas a polynomial in x or, more generally, as a sum of terms each of which containspowers of x <strong>and</strong> α separately, as opposed to a power of their sum or difference.To make the discussion general <strong>and</strong> the result applicable to a wide variety ofsituations, we will consider the general expansion of f(x) =(x + y) n ,wherex <strong>and</strong>y may st<strong>and</strong> <strong>for</strong> constants, variables or functions <strong>and</strong>, <strong>for</strong> the time being, n is apositive integer. It may not be obvious what <strong>for</strong>m the general expansion takesbut some idea can be obtained by carrying out the multiplication explicitly <strong>for</strong>small values of n. Thus we obtain successively(x + y) 1 = x + y,(x + y) 2 =(x + y)(x + y) =x 2 +2xy + y 2 ,(x + y) 3 =(x + y)(x 2 +2xy + y 2 )=x 3 +3x 2 y +3xy 2 + y 3 ,(x + y) 4 =(x + y)(x 3 +3x 2 y +3xy 2 + y 3 )=x 4 +4x 3 y +6x 2 y 2 +4xy 3 + y 4 .This does not establish a general <strong>for</strong>mula, but the regularity of the terms inthe expansions <strong>and</strong> the suggestion of a pattern in the coefficients indicate that ageneral <strong>for</strong>mula <strong>for</strong> power n will have n + 1 terms, that the powers of x <strong>and</strong> y inevery term will add up to n <strong>and</strong> that the coefficients of the first <strong>and</strong> last termswill be unity whilst those of the second <strong>and</strong> penultimate terms will be n.25


PRELIMINARY ALGEBRAIn fact, the general expression, the binomial expansion <strong>for</strong> power n, is given by∑k=n(x + y) n =n C k x n−k y k , (1.49)k=0where n C k is called the binomial coefficient <strong>and</strong> is expressed in terms of factorialfunctions by n!/[k!(n − k)!]. Clearly, simply to make such a statement does notconstitute proof of its validity, but, as we will see in subsection 1.5.2, (1.49) canbe proved using a method called induction. Be<strong>for</strong>e turning to that proof, weinvestigate some of the elementary properties of the binomial coefficients.1.5.1 Binomial coefficientsAs stated above, the binomial coefficients are defined by( )n n! nC k ≡k!(n − k)! ≡ <strong>for</strong> 0 ≤ k ≤ n, (1.50)kwhere in the second identity we give a common alternative notation <strong>for</strong> n C k .Obvious properties include(i) n C 0 = n C n =1,(ii) n C 1 = n C n−1 = n,(iii) n C k = n C n−k .We note that, <strong>for</strong> any given n, the largest coefficient in the binomial expansion isthe middle one (k = n/2) if n is even; the middle two coefficients (k = 1 2(n ± 1))are equal largest if n is odd. Somewhat less obvious is the resultn C k + n n!C k−1 =k!(n − k)! + n!(k − 1)!(n − k +1)!n![(n +1− k)+k]=k!(n +1− k)!(n +1)!=k!(n +1− k)! = n+1 C k . (1.51)An equivalent statement, in which k has been redefined as k +1,isn C k + n C k+1 = n+1 C k+1 . (1.52)1.5.2 Proof of the binomial expansionWe are now in a position to prove the binomial expansion (1.49). In doing so, weintroduce the reader to a procedure applicable to certain types of problems <strong>and</strong>known as the method of induction. The method is discussed much more fully insubsection 1.7.1.26


1.6 PROPERTIES OF BINOMIAL COEFFICIENTSWe start by assuming that (1.49) is true <strong>for</strong> some positive integer n = N.Wenowproceed to show that this implies that it must also be true <strong>for</strong> n = N+1, as follows:(x + y) N+1 =(x + y)k=0N∑N C k x N−k y kk=0N∑N∑=N C k x N+1−k y k +N C k x N−k y k+1=k=0N∑N+1∑N C k x N+1−k y k +N C j−1 x (N+1)−j y j ,k=0where in the first line we have used the assumption <strong>and</strong> in the third line havemoved the second summation index by unity, by writing k +1 = j. Wenowseparate off the first term of the first sum, N C 0 x N+1 , <strong>and</strong> write it as N+1 C 0 x N+1 ;we can do this since, as noted in (i) following (1.50), n C 0 =1<strong>for</strong>everyn. Similarly,the last term of the second summation can be replaced by N+1 C N+1 y N+1 .The remaining terms of each of the two summations are now written together,with the summation index denoted by k in both terms. Thus(x + y) N+1 = N+1 C 0 x N+1 += N+1 C 0 x N+1 +j=1N∑ ( NC k + N )C k−1 x (N+1)−k y k + N+1 C N+1 y N+1k=1N∑N+1 C k x (N+1)−k y k + N+1 C N+1 y N+1k=1N+1∑=N+1 C k x (N+1)−k y k .k=0In going from the first to the second line we have used result (1.51). Now weobserve that the final overall equation is just the original assumed result (1.49)but with n = N + 1. Thus it has been shown that if the binomial expansion isassumed to be true <strong>for</strong> n = N, thenitcanbeproved to be true <strong>for</strong> n = N +1. Butit holds trivially <strong>for</strong> n = 1, <strong>and</strong> there<strong>for</strong>e <strong>for</strong> n = 2 also. By the same token it isvalid <strong>for</strong> n =3, 4,..., <strong>and</strong> hence is established <strong>for</strong> all positive integers n.1.6 Properties of binomial coefficients1.6.1 Identities involving binomial coefficientsThere are many identities involving the binomial coefficients that can be deriveddirectly from their definition, <strong>and</strong> yet more that follow from their appearance inthe binomial expansion. Only the most elementary ones, given earlier, are worthcommitting to memory but, as illustrations, we now derive two results involvingsums of binomial coefficients.27


PRELIMINARY ALGEBRAThe first is a further application of the method of induction. Consider theproposal that, <strong>for</strong> any n ≥ 1<strong>and</strong>k ≥ 0,∑n−1k+s C k = n+k C k+1 . (1.53)s=0Notice that here n, the number of terms in the sum, is the parameter that varies,k is a fixed parameter, whilst s is a summation index <strong>and</strong> does not appear on theRHS of the equation.Now we suppose that the statement (1.53) about the value of the sum of thebinomial coefficients k C k , k+1 C k ,..., k+n−1 C k is true <strong>for</strong> n = N. Wenextwritedowna series with an extra term <strong>and</strong> determine the implications of the supposition <strong>for</strong>the new series:N+1−1∑N−1∑k+s C k =k+s C k + k+N C ks=0s=0= N+k C k+1 + N+k C k= N+k+1 C k+1 .But this is just proposal (1.53) with n now set equal to N + 1. To obtain the lastline, we have used (1.52), with n set equal to N + k.It only remains to consider the case n = 1, when the summation only containsone term <strong>and</strong> (1.53) reduces tok C k = 1+k C k+1 .This is trivially valid <strong>for</strong> any k since both sides are equal to unity, thus completingthe proof of (1.53) <strong>for</strong> all positive integers n.The second result, which gives a <strong>for</strong>mula <strong>for</strong> combining terms from two setsof binomial coefficients in a particular way (a kind of ‘convolution’, <strong>for</strong> readerswho are already familiar with this term), is derived by applying the binomialexpansion directly to the identity(x + y) p (x + y) q ≡ (x + y) p+q .Written in terms of binomial expansions, this readsp∑q∑∑p+qp C s x p−s y s q C t x q−t y t =p+q C r x p+q−r y r .s=0t=0We now equate coefficients of x p+q−r y r on the two sides of the equation, notingthat on the LHS all combinations of s <strong>and</strong> t such that s + t = r contribute. Thisgives as an identity thatr∑r∑p C q r−t C t = p+q C r =p C q t C r−t . (1.54)r=0t=0t=028


1.6 PROPERTIES OF BINOMIAL COEFFICIENTSWe have specifically included the second equality to emphasise the symmetricalnature of the relationship with respect to p <strong>and</strong> q.Further identities involving the coefficients can be obtained by giving x <strong>and</strong> yspecial values in the defining equation (1.49) <strong>for</strong> the expansion. If both are setequal to unity then we obtain (using the alternative notation so as to producefamiliarity with it)( ( ( ( n n n n+ + + ···+ =20)1)2)n)n , (1.55)whilst setting x = 1 <strong>and</strong> y = −1 yields( ( ( ( )n n n n− + − ···+(−1)0)1)2)n n=0. (1.56)1.6.2 Negative <strong>and</strong> non-integral values of nUp till now we have restricted n in the binomial expansion to be a positiveinteger. Negative values can be accommodated, but only at the cost of an infiniteseries of terms rather than the finite one represented by (1.49). For reasons thatare intuitively sensible <strong>and</strong> will be discussed in more detail in chapter 4, veryoften we require an expansion in which, at least ultimately, successive terms inthe infinite series decrease in magnitude. For this reason, if x>ywe consider(x + y) −m ,wherem itself is a positive integer, in the <strong>for</strong>m(x + y) n =(x + y) −m = x −m ( 1+ y x) −m.Since the ratio y/x is less than unity, terms containing higher powers of it will besmall in magnitude, whilst raising the unit term to any power will not affect itsmagnitude. If y>xthe roles of the two must be interchanged.We can now state, but will not explicitly prove, the <strong>for</strong>m of the binomialexpansion appropriate to negative values of n (n equal to −m):∑∞ ( y) k(x + y) n =(x + y) −m = x −m −m C k , (1.57)xwhere the hitherto undefined quantity −m C k , which appears to involve factorialsof negative numbers, is given by−m k m(m +1)···(m + k − 1) k (m + k − 1)!C k =(−1) =(−1) =(−1) k m+k−1 C k .k!(m − 1)!k!(1.58)The binomial coefficient on the extreme right of this equation has its normalmeaning <strong>and</strong> is well defined since m + k − 1 ≥ k.Thus we have a definition of binomial coefficients <strong>for</strong> negative integer valuesof n in terms of those <strong>for</strong> positive n. The connection between the two may not29k=0


PRELIMINARY ALGEBRAbe obvious, but they are both <strong>for</strong>med in the same way in terms of recurrencerelations. Whatever the sign of n, the series of coefficients n C k can be generatedby starting with n C 0 = 1 <strong>and</strong> using the recurrence relationn C k+1 = n − k n C k . (1.59)k +1The difference is that <strong>for</strong> positive integer n the series terminates when k = n,whereas <strong>for</strong> negative n there is no such termination – in line with the infiniteseries of terms in the corresponding expansion.Finally we note that, in fact, equation (1.59) generates the appropriate coefficients<strong>for</strong> all values of n, positive or negative, integer or non-integer, with theobvious exception of the case in which x = −y <strong>and</strong> n is negative. For non-integern the expansion does not terminate, even if n is positive.1.7 Some particular methods of proofMuch of the mathematics used by physicists <strong>and</strong> engineers is concerned withobtaining a particular value, <strong>for</strong>mula or function from a given set of data <strong>and</strong>stated conditions. However, just as it is essential in physics to <strong>for</strong>mulate the basiclaws <strong>and</strong> so be able to set boundaries on what can or cannot happen, so itis important in mathematics to be able to state general propositions about theoutcomes that are or are not possible. To this end one attempts to establishtheorems that state in as general a way as possible mathematical results thatapply to particular types of situation. We conclude this introductory chapter bydescribing two methods that can sometimes be used to prove particular classesof theorems.The two general methods of proof are known as proof by induction (whichhas already been met in this chapter) <strong>and</strong> proof by contradiction. They sharethe common characteristic that at an early stage in the proof an assumptionis made that a particular (unproven) statement is true; the consequences ofthat assumption are then explored. In an inductive proof the conclusion isreached that the assumption is self-consistent <strong>and</strong> has other equally consistentbut broader implications, which are then applied to establish the general validityof the assumption. A proof by contradiction, however, establishes an internalinconsistency <strong>and</strong> thus shows that the assumption is unsustainable; the naturalconsequence of this is that the negative of the assumption is established as true.Later in this book use will be made of these methods of proof to explore newterritory, e.g. to examine the properties of vector spaces, matrices <strong>and</strong> groups.However, at this stage we will draw our illustrative <strong>and</strong> test examples from earliersections of this chapter <strong>and</strong> other topics in elementary algebra <strong>and</strong> number theory.30


1.7 SOME PARTICULAR METHODS OF PROOF1.7.1 Proof by inductionThe proof of the binomial expansion given in subsection 1.5.2 <strong>and</strong> the identityestablished in subsection 1.6.1 have already shown the way in which an inductiveproof is carried through. They also indicated the main limitation of the method,namely that only an initially supposed result can be proved. Thus the methodof induction is of no use <strong>for</strong> deducing a previously unknown result; a putativeequation or result has to be arrived at by some other means, usually by noticingpatterns or by trial <strong>and</strong> error using simple values of the variables involved. Itwill also be clear that propositions that can be proved by induction are limitedto those containing a parameter that takes a range of integer values (usuallyinfinite).For a proposition involving a parameter n, the five steps in a proof usinginduction are as follows.(i) Formulate the supposed result <strong>for</strong> general n.(ii) Suppose (i) to be true <strong>for</strong> n = N (or more generally <strong>for</strong> all values ofn ≤ N; seebelow),whereN is restricted to lie in the stated range.(iii) Show, using only proven results <strong>and</strong> supposition (ii), that proposition (i)is true <strong>for</strong> n = N +1.(iv) Demonstrate directly, <strong>and</strong> without any assumptions, that proposition (i) istrue when n takes the lowest value in its range.(v) It then follows from (iii) <strong>and</strong> (iv) that the proposition is valid <strong>for</strong> all valuesof n in the stated range.(It should be noted that, although many proofs at stage (iii) require the validityof the proposition only <strong>for</strong> n = N, some require it <strong>for</strong> all n less than or equal to N– hence the <strong>for</strong>m of inequality given in parentheses in the stage (ii) assumption.)To illustrate further the method of induction, we now apply it to two workedexamples; the first concerns the sum of the squares of the first n natural numbers.◮Prove that the sum of the squares of the first n natural numbers is given byn∑r 2 = 1 n(n + 1)(2n +1). (1.60)6r=1As previously we start by assuming the result is true <strong>for</strong> n = N. Then it follows that∑N+1N∑r 2 = r 2 +(N +1) 2r=1r=1= 1 N(N + 1)(2N +1)+(N +1)26= 1 (N +1)[N(2N +1)+6N +6]6= 1 (N + 1)[(2N +3)(N +2)]6= 1 (N +1)[(N +1)+1][2(N +1)+1].631


PRELIMINARY ALGEBRAThis is precisely the original assumption, but with N replaced by N + 1. To complete theproof we only have to verify (1.60) <strong>for</strong> n = 1. This is trivially done <strong>and</strong> establishes theresult <strong>for</strong> all positive n. The same <strong>and</strong> related results are obtained by a different methodin subsection 4.2.5. ◭Our second example is somewhat more complex <strong>and</strong> involves two nested proofsby induction: whilst trying to establish the main result by induction, we find thatwe are faced with a second proposition which itself requires an inductive proof.◮Show that Q(n) =n 4 +2n 3 +2n 2 + n is divisible by 6 (without remainder) <strong>for</strong> all positiveinteger values of n.Again we start by assuming the result is true <strong>for</strong> some particular value N of n, whilstnoting that it is trivially true <strong>for</strong> n = 0. We next examine Q(N + 1), writing each of itsterms as a binomial expansion:Q(N +1)=(N +1) 4 +2(N +1) 3 +2(N +1) 2 +(N +1)=(N 4 +4N 3 +6N 2 +4N +1)+2(N 3 +3N 2 +3N +1)+2(N 2 +2N +1)+(N +1)=(N 4 +2N 3 +2N 2 + N)+(4N 3 +12N 2 +14N +6).Now, by our assumption, the group of terms within the first parentheses in the last lineis divisible by 6 <strong>and</strong> clearly so are the terms 12N 2 <strong>and</strong> 6 within the second parentheses.Thus it comes down to deciding whether 4N 3 +14N is divisible by 6 – or equivalently,whether R(N) =2N 3 +7N is divisible by 3.To settle this latter question we try using a second inductive proof <strong>and</strong> assume thatR(N) is divisible by 3 <strong>for</strong> N = M, whilst again noting that the proposition is trivially true<strong>for</strong> N = M = 0. This time we examine R(M +1):R(M +1)=2(M +1) 3 +7(M +1)=2(M 3 +3M 2 +3M +1)+7(M +1)=(2M 3 +7M)+3(2M 2 +2M +3)By assumption, the first group of terms in the last line is divisible by 3 <strong>and</strong> the secondgroup is patently so. We thus conclude that R(N) is divisible by 3 <strong>for</strong> all N ≥ M, <strong>and</strong>taking M = 0 shows that it is divisible by 3 <strong>for</strong> all N.We can now return to the main proposition <strong>and</strong> conclude that since R(N) =2N 3 +7Nis divisible by 3, 4N 3 +12N 2 +14N + 6 is divisible by 6. This in turn establishes that thedivisibility of Q(N + 1) by 6 follows from the assumption that Q(N) divides by 6. SinceQ(0) clearly divides by 6, the proposition in the question is established <strong>for</strong> all values of n. ◭1.7.2 Proof by contradictionThe second general line of proof, but again one that is normally only useful whenthe result is already suspected, is proof by contradiction. The questions it canattempt to answer are only those that can be expressed in a proposition thatis either true or false. Clearly, it could be argued that any mathematical resultcan be so expressed but, if the proposition is no more than a guess, the chancesof success are negligible. Valid propositions containing even modest <strong>for</strong>mulaeare either the result of true inspiration or, much more normally, yet anotherreworking of an old chestnut!32


1.7 SOME PARTICULAR METHODS OF PROOFThe essence of the method is to exploit the fact that mathematics is requiredto be self-consistent, so that, <strong>for</strong> example, two calculations of the same quantity,starting from the same given data but proceeding by different methods, must givethe same answer. Equally, it must not be possible to follow a line of reasoning <strong>and</strong>draw a conclusion that contradicts either the input data or any other conclusionbased upon the same data.It is this requirement on which the method of proof by contradiction is based.The crux of the method is to assume that the proposition to be proved isnot true, <strong>and</strong> then use this incorrect assumption <strong>and</strong> ‘watertight’ reasoning todraw a conclusion that contradicts the assumption. The only way out of theself-contradiction is then to conclude that the assumption was indeed false <strong>and</strong>there<strong>for</strong>e that the proposition is true.It must be emphasised that once a (false) contrary assumption has been made,every subsequent conclusion in the argument must follow of necessity. Proof bycontradiction fails if at any stage we have to admit ‘this may or may not bethe case’. That is, each step in the argument must be a necessary consequence ofresults that precede it (taken together with the assumption), rather than simply apossible consequence.It should also be added that if no contradiction can be found using soundreasoning based on the assumption then no conclusion can be drawn about eitherthe proposition or its negative <strong>and</strong> some other approach must be tried.We illustrate the general method with an example in which the mathematicalreasoning is straight<strong>for</strong>ward, so that attention can be focussed on the structureof the proof.◮A rational number r is a fraction r = p/q in which p <strong>and</strong> q are integers with q positive.Further, r is expressed in its lowest terms, any integer common factor of p <strong>and</strong> q havingbeen divided out.Prove that the square root of an integer m cannot be a rational number, unless the squareroot itself is an integer.We begin by supposing that the stated result is not true <strong>and</strong> that we can write an equation√ m = r =pq<strong>for</strong> integers m, p, q with q ≠1.It then follows that p 2 = mq 2 .But,sincer is expressed in its lowest terms, p <strong>and</strong> q, <strong>and</strong>hence p 2 <strong>and</strong> q 2 , have no factors in common. However, m is an integer; this is only possibleif q =1<strong>and</strong>p 2 = m. This conclusion contradicts the requirement that q ≠1<strong>and</strong>soleadsto the conclusion that it was wrong to suppose that √ m can be expressed as a non-integerrational number. This completes the proof of the statement in the question. ◭Our second worked example, also taken from elementary number theory,involves slightly more complicated mathematical reasoning but again exhibits thestructure associated with this type of proof.33


PRELIMINARY ALGEBRA◮The prime integers p i are labelled in ascending order, thus p 1 =1, p 2 =2, p 5 =7, etc.Show that there is no largest prime number.Assume, on the contrary, that there is a largest prime <strong>and</strong> let it be p N .Considernowthenumber q <strong>for</strong>med by multiplying together all the primes from p 1 to p N <strong>and</strong> then addingone to the product, i.e.q = p 1 p 2 ···p N +1.By our assumption p N is the largest prime, <strong>and</strong> so no number can have a prime factorgreater than this. However, <strong>for</strong> every prime p i , i =1, 2,...,N, the quotient q/p i has the<strong>for</strong>m M i +(1/p i )withM i an integer <strong>and</strong> 1/p i non-integer. This means that q/p i cannot bean integer <strong>and</strong> so p i cannot be a divisor of q.Since q is not divisible by any of the (assumed) finite set of primes, it must be itself aprime. As q is also clearly greater than p N , we have a contradiction. This shows that ourassumption that there is a largest prime integer must be false, <strong>and</strong> so it follows that thereis no largest prime integer.It should be noted that the given construction <strong>for</strong> q does not generate all the primesthat actually exist (e.g. <strong>for</strong> N =3,q = 7 rather than the next actual prime value of 5, isfound), but this does not matter <strong>for</strong> the purposes of our proof by contradiction. ◭1.7.3 Necessary <strong>and</strong> sufficient conditionsAs the final topic in this introductory chapter, we consider briefly the notionof, <strong>and</strong> distinction between, necessary <strong>and</strong> sufficient conditions in the contextof proving a mathematical proposition. In ordinary English the distinction iswell defined, <strong>and</strong> that distinction is maintained in mathematics. However, inthe authors’ experience students tend to overlook it <strong>and</strong> assume (wrongly) that,having proved that the validity of proposition A implies the truth of propositionB, it follows by ‘reversing the argument’ that the validity of B automaticallyimplies that of A.As an example, let proposition A be that an integer N is divisible withoutremainder by 6, <strong>and</strong> proposition B be that N is divisible without remainder by2. Clearly, if A is true then it follows that B is true, i.e. A is a sufficient condition<strong>for</strong> B; it is not however a necessary condition, as is trivially shown by taking Nas 8. Conversely, the same value of N shows that whilst the validity of B is anecessary condition <strong>for</strong> A to hold, it is not sufficient.An alternative terminology to ‘necessary’ <strong>and</strong> ‘sufficient’ often employed bymathematicians is that of ‘if’ <strong>and</strong> ‘only if’, particularly in the combination ‘if <strong>and</strong>only if’ which is usually written as IFF or denoted by a double-headed arrow⇐⇒ . The equivalent statements can be summarised byA if B A is true if B is true or B =⇒ A,B is a sufficient condition <strong>for</strong> A B =⇒ A,A only if B A is true only if B is true or A =⇒ B,B is a necessary consequence of A A =⇒ B,34


1.7 SOME PARTICULAR METHODS OF PROOFA IFF B A is true if <strong>and</strong> only if B is true or B ⇐⇒ A,A <strong>and</strong> B necessarily imply each other B ⇐⇒ A.Although at this stage in the book we are able to employ <strong>for</strong> illustrative purposesonly simple <strong>and</strong> fairly obvious results, the following example is given as a modelof how necessary <strong>and</strong> sufficient conditions should be proved. The essential pointis that <strong>for</strong> the second part of the proof (whether it be the ‘necessary’ part or the‘sufficient’ part) one needs to start again from scratch; more often than not, thelines of the second part of the proof will not be simply those of the first writtenin reverse order.◮Prove that (A) a function f(x) is a quadratic polynomial with zeros at x =2<strong>and</strong> x =3if <strong>and</strong> only if (B) the function f(x) has the <strong>for</strong>m λ(x 2 − 5x +6) with λ a non-zero constant.(1) Assume A, i.e.thatf(x) is a quadratic polynomial with zeros at x =2<strong>and</strong>x =3.Letits <strong>for</strong>m be ax 2 + bx + c with a ≠ 0. Then we have4a +2b + c =0,9a +3b + c =0,<strong>and</strong> subtraction shows that 5a + b =0<strong>and</strong>b = −5a. Substitution of this into the first ofthe above equations gives c = −4a − 2b = −4a +10a =6a. Thus, it follows thatf(x) =a(x 2 − 5x +6) with a ≠0,<strong>and</strong> establishes the ‘A only if B’ part of the stated result.(2) Now assume that f(x) has the <strong>for</strong>m λ(x 2 − 5x +6)withλ a non-zero constant. Firstlywe note that f(x) is a quadratic polynomial, <strong>and</strong> so it only remains to prove that itszeros occur at x =2<strong>and</strong>x =3.Considerf(x) = 0, which, after dividing through by thenon-zero constant λ, givesx 2 − 5x +6=0.We proceed by using a technique known as completing the square, <strong>for</strong> the purposes ofillustration, although the factorisation of the above equation should be clear to the reader.Thus we writex 2 − 5x +( 5 2 )2 − ( 5 2 )2 +6=0,(x − 5 2 )2 = 1 , 4x − 5 = ± 1 .2 2The two roots of f(x) = 0 are there<strong>for</strong>e x =2<strong>and</strong>x =3;thesex-values give the zerosof f(x). This establishes the second (‘A if B’) part of the result. Thus we have shownthat the assumption of either condition implies the validity of the other <strong>and</strong> the proof iscomplete. ◭It should be noted that the propositions have to be carefully <strong>and</strong> precisely<strong>for</strong>mulated. If, <strong>for</strong> example, the word ‘quadratic’ were omitted from A, statementB would still be a sufficient condition <strong>for</strong> A but not a necessary one, since f(x)could then be x 3 − 4x 2 + x + 6 <strong>and</strong> A would not require B. Omitting the constantλ from the stated <strong>for</strong>m of f(x) inB has the same effect. Conversely, if A were tostate that f(x) =3(x − 2)(x − 3) then B would be a necessary condition <strong>for</strong> A butnot a sufficient one.35


PRELIMINARY ALGEBRA1.8 ExercisesPolynomial equations1.1 Continue the investigation of equation (1.7), namelyg(x) =4x 3 +3x 2 − 6x − 1,as follows.(a) Make a table of values of g(x) <strong>for</strong> integer values of x between −2 <strong>and</strong>2.Useit <strong>and</strong> the in<strong>for</strong>mation derived in the text to draw a graph <strong>and</strong> so determinethe roots of g(x) = 0 as accurately as possible.(b) Find one accurate root of g(x) = 0 by inspection <strong>and</strong> hence determine precisevalues <strong>for</strong> the other two roots.(c) Show that f(x) =4x 3 +3x 2 − 6x − k = 0 has only one real root unless−5 ≤ k ≤ 7 . 41.2 Determine how the number of real roots of the equationg(x) =4x 3 − 17x 2 +10x + k =0depends upon k. Are there any cases <strong>for</strong> which the equation has exactly twodistinct real roots?1.3 Continue the analysis of the polynomial equationf(x) =x 7 +5x 6 + x 4 − x 3 + x 2 − 2=0,investigated in subsection 1.1.1, as follows.(a) By writing the fifth-degree polynomial appearing in the expression <strong>for</strong> f ′ (x)in the <strong>for</strong>m 7x 5 +30x 4 + a(x − b) 2 + c, show that there is in fact only onepositive root of f(x) =0.(b) By evaluating f(1), f(0) <strong>and</strong> f(−1), <strong>and</strong> by inspecting the <strong>for</strong>m of f(x) <strong>for</strong>negative values of x, determine what you can about the positions of the realroots of f(x) =0.1.4 Given that x =2isonerootofg(x) =2x 4 +4x 3 − 9x 2 − 11x − 6=0,use factorisation to determine how many real roots it has.1.5 Construct the quadratic equations that have the following pairs of roots:(a) −6, −3; (b) 0, 4; (c) 2, 2;(d)3+2i, 3 − 2i, wherei 2 = −1.1.6 Use the results of (i) equation (1.13), (ii) equation (1.12) <strong>and</strong> (iii) equation (1.14)to prove that if the roots of 3x 3 − x 2 − 10x +8=0areα 1 ,α 2 <strong>and</strong> α 3 then(a) α1 −1 + α −12+ α −13=5/4,(b) α 2 1 + α2 2 + α2 3 =61/9,(c) α 3 1 + α3 2 + α3 3 = −125/27.(d) Convince yourself that eliminating (say) α 2 <strong>and</strong> α 3 from (i), (ii) <strong>and</strong> (iii) doesnot give a simple explicit way of finding α 1 .1.7 Prove thatby consideringTrigonometric identitiescos π 12 = √3+12 √ 236


1.8 EXERCISES(a) the sum of the sines of π/3 <strong>and</strong>π/6,(b) the sine of the sum of π/3 <strong>and</strong>π/4.1.8 The following exercises are based on the half-angle <strong>for</strong>mulae.(a) Use the fact that sin(π/6) = 1/2 to prove that tan(π/12) = 2 − √ 3.(b) Use the result of (a) to show further that tan(π/24) = q(2 − q) whereq 2 =2+ √ 3.1.9 Find the real solutions of(a) 3 sin θ − 4cosθ =2,(b) 4 sin θ +3cosθ =6,(c) 12 sin θ − 5cosθ = −6.1.10 If s = sin(π/8), prove that8s 4 − 8s 2 +1=0,<strong>and</strong> hence show that s =[(2− √ 2)/4] 1/2 .1.11 Find all the solutions ofsin θ +sin4θ =sin2θ +sin3θthat lie in the range −π


PRELIMINARY ALGEBRA1.16 Express the following in partial fraction <strong>for</strong>m:(a) 2x3 − 5x +1x 2 − 2x − 8 , (b) x2 + x − 1x 2 + x − 2 .1.17 Rearrange the following functions in partial fraction <strong>for</strong>m:x − 6(a)x 3 − x 2 +4x − 4 , (b) x3 +3x 2 + x +19.x 4 +10x 2 +91.18 Resolve the following into partial fractions in such a way that x does not appearin any numerator:(a)2x 2 + x +1(x − 1) 2 (x +3) , (b) x 2 − 2x 3 +8x 2 +16x , (c) x 3 − x − 1(x +3) 3 (x +1) .Binomial expansion1.19 Evaluate those of the following that are defined: (a) 5 C 3 ,(b) 3 C 5 ,(c) −5 C 3 ,(d)−3 C 5 .1.20 Use a binomial expansion to evaluate 1/ √ 4.2 to five places of decimals, <strong>and</strong>compare it with the accurate answer obtained using a calculator.Proof by induction <strong>and</strong> contradiction1.21 Prove by induction thatn∑n∑r = 1 n(n +1) <strong>and</strong> r 3 = 1 2 4 n2 (n +1) 2 .r=11.22 Prove by induction that1+r + r 2 + ···+ r k + ···+ r n = 1 − rn+1.1 − r1.23 Prove that 3 2n +7,wheren is a non-negative integer, is divisible by 8.1.24 If a sequence of terms, u n , satisfies the recurrence relation u n+1 =(1− x)u n + nx,with u 1 = 0, show, by induction, that, <strong>for</strong> n ≥ 1,r=1u n = 1 x [nx − 1+(1− x)n ].1.25 Prove by induction thatn∑( )1 θ2 tan = 1 ( ) θ r 2 r 2 cot − cot θ.n 2 n r=11.26 The quantities a i in this exercise are all positive real numbers.(a) Show that( ) 2 a1 + a 2a 1 a 2 ≤.2(b) Hence prove, by induction on m, that( ) p a1 + a 2 + ···+ a pa 1 a 2 ···a p ≤,pwhere p =2 m with m a positive integer. Note that each increase of m byunity doubles the number of factors in the product.38


1.9 HINTS AND ANSWERS1.27 Establish the values of k <strong>for</strong> which the binomial coefficient p C k is divisible by pwhen p is a prime number. Use your result <strong>and</strong> the method of induction to provethat n p − n is divisible by p <strong>for</strong> all integers n <strong>and</strong> all prime numbers p. Deducethat n 5 − n isdivisibleby30<strong>for</strong>anyintegern.1.28 An arithmetic progression of integers a n is one in which a n = a 0 + nd, wherea 0<strong>and</strong> d are integers <strong>and</strong> n takes successive values 0, 1, 2,....(a) Show that if any one term of the progression is the cube of an integer thenso are infinitely many others.(b) Show that no cube of an integer can be expressed as 7n + 5 <strong>for</strong> some positiveinteger n.1.29 Prove, by the method of contradiction, that the equationx n + a n−1 x n−1 + ···+ a 1 x + a 0 =0,in which all the coefficients a i are integers, cannot have a rational root, unlessthat root is an integer. Deduce that any integral root must be a divisor of a 0 <strong>and</strong>hence find all rational roots of(a) x 4 +6x 3 +4x 2 +5x +4=0,(b) x 4 +5x 3 +2x 2 − 10x +6=0.Necessary <strong>and</strong> sufficient conditions1.30 Prove that the equation ax 2 + bx + c =0,inwhicha, b <strong>and</strong> c are real <strong>and</strong> a>0,has two real distinct solutions IFF b 2 > 4ac.1.31 For the real variable x, show that a sufficient, but not necessary, condition <strong>for</strong>f(x) =x(x + 1)(2x + 1) to be divisible by 6 is that x is an integer.1.32 Given that at least one of a <strong>and</strong> b, <strong>and</strong> at least one of c <strong>and</strong> d, are non-zero,show that ad = bc is both a necessary <strong>and</strong> sufficient condition <strong>for</strong> the equationsax + by =0,cx + dy =0,to have a solution in which at least one of x <strong>and</strong> y is non-zero.1.33 The coefficients a i in the polynomial Q(x) =a 4 x 4 + a 3 x 3 + a 2 x 2 + a 1 x are allintegers. Show that Q(n) is divisible by 24 <strong>for</strong> all integers n ≥ 0 if <strong>and</strong> only if allof the following conditions are satisfied:(i) 2a 4 + a 3 is divisible by 4;(ii) a 4 + a 2 is divisible by 12;(iii) a 4 + a 3 + a 2 + a 1 is divisible by 24.1.9 Hints <strong>and</strong> answers1.1 (b) The roots are 1, 1 (−7+√ 33) = −0.1569, 1 (−7 − √ 33) = −1.593. (c) −5 <strong>and</strong>8 87are the values of k that make f(−1) <strong>and</strong> f( 1 ) equal to zero.4 21.3 (a) a =4,b= 3 23<strong>and</strong> c = are all positive. There<strong>for</strong>e 8 16 f′ (x) > 0<strong>for</strong>allx>0.(b) f(1) = 5, f(0) = −2 <strong>and</strong>f(−1) = 5, <strong>and</strong> so there is at least one root in eachof the ranges 0


PRELIMINARY ALGEBRA1.11 Show that the equation is equivalent to sin(5θ/2) sin(θ)sin(θ/2) = 0.Solutions are −4π/5, −2π/5, 0, 2π/5, 4π/5,π.Thesolutionθ = 0 has multiplicity3.1.13 (a) A circle of radius 5 centred on (−3, −4).(b) A hyperbola with ‘centre’ (3, −2) <strong>and</strong> ‘semi-axes’ 2 <strong>and</strong> 3.(c) The expression factorises into two lines, x +2y − 3=0<strong>and</strong>2x + y +2=0.(d) Write the expression as (x+y) 2 =8(x−y) to see that it represents a parabolapassing through the origin, with the line x + y = 0 as its axis of symmetry.51.15 (a)7(x − 2) + 97(x +5) , (b) − 43x + 43(x − 3) .1.17 (a) x +2x 2 +4 − 1x +1, (b)x − 1 x 2 +9 + 2x 2 +1 .1.19 (a) 10, (b) not defined, (c) −35, (d) −21.1.21 Look <strong>for</strong> factors common to the n = N sum <strong>and</strong> the additional n = N +1term,so as to reduce the sum <strong>for</strong> n = N +1toasingleterm.1.23 Write 3 2n as 8m − 7.1.25 Use the half-angle <strong>for</strong>mulae of equations (1.32) to (1.34) to relate functions ofθ/2 k to those of θ/2 k+1 .1.27 Divisible <strong>for</strong> k =1, 2,...,p− 1. Exp<strong>and</strong> (n +1) p as n p + ∑ p−11p C k n k + 1. Applythe stated result <strong>for</strong> p = 5. Note that n 5 − n = n(n − 1)(n +1)(n 2 + 1); the productof any three consecutive integers must divide by both 2 <strong>and</strong> 3.1.29 By assuming x = p/q with q ≠ 1, show that a fraction −p n /q isequaltoaninteger a n−1 p n−1 + ···+ a 1 pq n−2 + a 0 q n−1 . This is a contradiction, <strong>and</strong> is onlyresolved if q = 1 <strong>and</strong> the root is an integer.(a) The only possible c<strong>and</strong>idates are ±1, ±2, ±4. None is a root.(b) The only possible c<strong>and</strong>idates are ±1, ±2, ±3, ±6. Only −3 isaroot.1.31 f(x) can be written as x(x +1)(x +2)+x(x +1)(x − 1). Each term consists ofthe product of three consecutive integers, of which one must there<strong>for</strong>e divide by2 <strong>and</strong> (a different) one by 3. Thus each term separately divides by 6, <strong>and</strong> sothere<strong>for</strong>e does f(x). Note that if x is the root of 2x 3 +3x 2 + x − 24 = 0 that liesnear the non-integer value x =1.826, then x(x + 1)(2x + 1) = 24 <strong>and</strong> there<strong>for</strong>edivides by 6.1.33 Note that, e.g., the condition <strong>for</strong> 6a 4 + a 3 to be divisible by 4 is the same as thecondition <strong>for</strong> 2a 4 + a 3 to be divisible by 4.For the necessary (only if) part of the proof set n =1, 2, 3 <strong>and</strong> take integercombinations of the resulting equations.For the sufficient (if) part of the proof use the stated conditions to prove theproposition by induction. Note that n 3 − n is divisible by 6 <strong>and</strong> that n 2 + n iseven.40


2Preliminary calculusThis chapter is concerned with the <strong>for</strong>malism of probably the most widely usedmathematical technique in the physical sciences, namely the calculus. The chapterdivides into two sections. The first deals with the process of differentiation <strong>and</strong> thesecond with its inverse process, integration. The material covered is essential <strong>for</strong>the remainder of the book <strong>and</strong> serves as a reference. Readers who have previouslystudied these topics should ensure familiarity by looking at the worked examplesin the main text <strong>and</strong> by attempting the exercises at the end of the chapter.2.1 DifferentiationDifferentiation is the process of determining how quickly or slowly a functionvaries, as the quantity on which it depends, its argument, is changed. Morespecifically it is the procedure <strong>for</strong> obtaining an expression (numerical or algebraic)<strong>for</strong> the rate of change of the function with respect to its argument. Familiarexamples of rates of change include acceleration (the rate of change of velocity)<strong>and</strong> chemical reaction rate (the rate of change of chemical composition). Bothacceleration <strong>and</strong> reaction rate give a measure of the change of a quantity withrespect to time. However, differentiation may also be applied to changes withrespect to other quantities, <strong>for</strong> example the change in pressure with respect to achange in temperature.Although it will not be apparent from what we have said so far, differentiationis in fact a limiting process, that is, it deals only with the infinitesimal change inone quantity resulting from an infinitesimal change in another.2.1.1 Differentiation from first principlesLet us consider a function f(x) that depends on only one variable x, together withnumerical constants, <strong>for</strong> example, f(x) =3x 2 or f(x) =sinx or f(x) =2+3/x.41


PRELIMINARY CALCULUSf(x +∆x)AP∆ff(x)θx∆xx +∆xFigure 2.1 The graph of a function f(x) showing that the gradient or slopeof the function at P ,givenbytanθ, is approximately equal to ∆f/∆x.Figure 2.1 shows an example of such a function. Near any particular point,P , the value of the function changes by an amount ∆f, say, as x changesby a small amount ∆x. The slope of the tangent to the graph of f(x) at Pis then approximately ∆f/∆x, <strong>and</strong> the change in the value of the function is∆f = f(x +∆x) − f(x). In order to calculate the true value of the gradient, orfirst derivative, of the function at P , we must let ∆x become infinitesimally small.We there<strong>for</strong>e define the first derivative of f(x) asf ′ (x) ≡ df(x)dx≡ lim∆x→0f(x +∆x) − f(x), (2.1)∆xprovided that the limit exists. The limit will depend in almost all cases on thevalue of x. If the limit does exist at a point x = a then the function is said to bedifferentiable at a; otherwise it is said to be non-differentiable at a. The<strong>for</strong>malconcept of a limit <strong>and</strong> its existence or non-existence is discussed in chapter 4; <strong>for</strong>present purposes we will adopt an intuitive approach.In the definition (2.1), we allow ∆x to tend to zero from either positive ornegative values <strong>and</strong> require the same limit to be obtained in both cases. Afunction that is differentiable at a is necessarily continuous at a (there must beno jump in the value of the function at a), though the converse is not necessarilytrue. This latter assertion is illustrated in figure 2.1: the function is continuousat the ‘kink’ A but the two limits of the gradient as ∆x tends to zero frompositive or negative values are different <strong>and</strong> so the function is not differentiableat A.It should be clear from the above discussion that near the point P we may42


2.1 DIFFERENTIATIONapproximate the change in the value of the function, ∆f, that results from a smallchange ∆x in x by∆f ≈ df(x) ∆x. (2.2)dxAs one would expect, the approximation improves as the value of ∆x is reduced.In the limit in which the change ∆x becomes infinitesimally small, we denote itby the differential dx, <strong>and</strong> (2.2) readsdf = df(x) dx. (2.3)dxThis equality relates the infinitesimal change in the function, df, to the infinitesimalchange dx that causes it.So far we have discussed only the first derivative of a function. However, wecan also define the second derivative as the gradient of the gradient of a function.Again we use the definition (2.1) but now with f(x) replaced by f ′ (x). Hence thesecond derivative is defined byf ′′ f ′ (x +∆x) − f ′ (x)(x) ≡ lim, (2.4)∆x→0 ∆xprovided that the limit exists. A physical example of a second derivative is thesecond derivative of the distance travelled by a particle with respect to time. Sincethe first derivative of distance travelled gives the particle’s velocity, the secondderivative gives its acceleration.We can continue in this manner, the nth derivative of the function f(x) beingdefined byf (n) f (n−1) (x +∆x) − f (n−1) (x)(x) ≡ lim. (2.5)∆x→0 ∆xIt should be noted that with this notation f ′ (x) ≡ f (1) (x), f ′′ (x) ≡ f (2) (x), etc., <strong>and</strong>that <strong>for</strong>mally f (0) (x) ≡ f(x).All this should be familiar to the reader, though perhaps not with such <strong>for</strong>maldefinitions. The following example shows the differentiation of f(x) =x 2 from firstprinciples. In practice, however, it is desirable simply to remember the derivativesof st<strong>and</strong>ard functions; the techniques given in the remainder of this section canbe applied to find more complicated derivatives.43


PRELIMINARY CALCULUS◮Find from first principles the derivative with respect to x of f(x) =x 2 .Using the definition (2.1),f ′ f(x +∆x) − f(x)(x) = lim∆x→0 ∆x(x +∆x) 2 − x 2= lim∆x→0 ∆x2x∆x +(∆x) 2= lim∆x→0 ∆x= lim (2x +∆x).∆x→0As ∆x tends to zero, 2x +∆x tends towards 2x, hencef ′ (x) =2x. ◭Derivatives of other functions can be obtained in the same way. The derivativesof some simple functions are listed below (note that a is a constant):ddx (xn ) = nx n−1 ,dd(sin ax) = a cos ax,dxddx (tan ax) = a sec2 ax,ddx (eax ) = ae ax ,d(cos ax) = −a sin ax,dxddx (cot ax) = −a cosec2 ax,d(cos −1 x )−1= √dx a a2 − x ,2ddx (ln ax) = 1 x ,(sec ax) = a sec ax tan ax,dxd(cosec ax) = −a cosec ax cot ax,dxd(sin −1 x )1= √dx a a2 − x ,2d(tan −1 x )a=dx a a 2 + x 2 .Differentiation from first principles emphasises the definition of a derivative asthe gradient of a function. However, <strong>for</strong> most practical purposes, returning to thedefinition (2.1) is time consuming <strong>and</strong> does not aid our underst<strong>and</strong>ing. Instead, asmentioned above, we employ a number of techniques, which use the derivativeslisted above as ‘building blocks’, to evaluate the derivatives of more complicatedfunctions than hitherto encountered. Subsections 2.1.2–2.1.7 develop the methodsrequired.2.1.2 Differentiation of productsAs a first example of the differentiation of a more complicated function, weconsider finding the derivative of a function f(x) that can be written as theproduct of two other functions of x, namely f(x) =u(x)v(x). For example, iff(x) = x 3 sin x then we might take u(x) = x 3 <strong>and</strong> v(x) = sinx. Clearly the44


2.1 DIFFERENTIATIONseparation is not unique. (In the given example, possible alternative break-upswould be u(x) =x 2 , v(x) =x sin x, orevenu(x) =x 4 tan x, v(x) =x −1 cos x.)The purpose of the separation is to split the function into two (or more) parts,of which we know the derivatives (or at least we can evaluate these derivativesmore easily than that of the whole). We would gain little, however, if we didnot know the relationship between the derivative of f <strong>and</strong> those of u <strong>and</strong> v.Fortunately, they are very simply related, as we shall now show.Since f(x) is written as the product u(x)v(x), it follows thatf(x +∆x) − f(x) =u(x +∆x)v(x +∆x) − u(x)v(x)= u(x +∆x)[v(x +∆x) − v(x)] + [u(x +∆x) − u(x)]v(x).From the definition of a derivative (2.1),dfdx = lim f(x +∆x) − f(x)∆x→0{∆x= lim u(x +∆x)∆x→0 ∆x[ v(x +∆x) − v(x)]+[ u(x +∆x) − u(x)∆x] }v(x) .In the limit ∆x → 0, the factors in square brackets become dv/dx <strong>and</strong> du/dx(by the definitions of these quantities) <strong>and</strong> u(x +∆x) simply becomes u(x).Consequently we obtaindfdx = d [u(x)v(x)] = u(x)dv(x)dx dx+ du(x) v(x). (2.6)dxIn primed notation <strong>and</strong> without writing the argument x explicitly, (2.6) is statedconcisely asf ′ =(uv) ′ = uv ′ + u ′ v. (2.7)This is a general result obtained without making any assumptions about thespecific <strong>for</strong>ms f, u <strong>and</strong> v, other than that f(x) =u(x)v(x). In words, the resultreads as follows. The derivative of the product of two functions is equal to thefirst function times the derivative of the second plus the second function times thederivative of the first.◮Find the derivative with respect to x of f(x) =x 3 sin x.Using the product rule, (2.6),ddx (x3 sin x) =x 3 d d(sin x)+dx dx (x3 )sinx= x 3 cos x +3x 2 sin x. ◭The product rule may readily be extended to the product of three or morefunctions. Considering the functionf(x) =u(x)v(x)w(x) (2.8)45


PRELIMINARY CALCULUS<strong>and</strong> using (2.6), we obtain, as be<strong>for</strong>e omitting the argument,dfdx = u d du(vw) +dx dx vw.Using (2.6) again to exp<strong>and</strong> the first term on the RHS gives the complete resultddw(uvw) =uvdx dx + u dvdx w + du vw (2.9)dxor(uvw) ′ = uvw ′ + uv ′ w + u ′ vw. (2.10)It is readily apparent that this can be extended to products containing any numbern of factors; the expression <strong>for</strong> the derivative will then consist of n terms withthe prime appearing in successive terms on each of the n factors in turn. This isprobably the easiest way to recall the product rule.2.1.3 The chain ruleProducts are just one type of complicated function that we may encounter indifferentiation. Another is the function of a function, e.g. f(x) =(3+x 2 ) 3 = u(x) 3 ,where u(x) =3+x 2 .If∆f, ∆u <strong>and</strong> ∆x are small finite quantities, it follows that∆f∆x = ∆f ∆u∆u ∆x ;As the quantities become infinitesimally small we obtaindfdx = df dudu dx . (2.11)This is the chain rule, which we must apply when differentiating a function of afunction.◮Find the derivative with respect to x of f(x) =(3+x 2 ) 3 .Rewriting the function as f(x) =u 3 ,whereu(x) =3+x 2 , <strong>and</strong> applying (2.11) we finddf du d=3u2dx dx =3u2 dx (3 + x2 )=3u 2 × 2x =6x(3 + x 2 ) 2 . ◭Similarly, the derivative with respect to x of f(x) =1/v(x) may be obtained byrewriting the function as f(x) =v −1 <strong>and</strong> applying (2.11):df dv= −v−2dx dx = − 1 dvv 2 dx . (2.12)The chain rule is also useful <strong>for</strong> calculating the derivative of a function f withrespect to x when both x <strong>and</strong> f are written in terms of a variable (or parameter),say t.46


2.1 DIFFERENTIATION◮Find the derivative with respect to x of f(t) =2at, wherex = at 2 .We could of course substitute <strong>for</strong> t <strong>and</strong> then differentiate f as a function of x, but in thiscase it is quicker to usedfdx = df dtdt dx =2a 12at = 1 t ,where we have used the fact that( ) −1dt dxdx = . ◭dt2.1.4 Differentiation of quotientsApplying (2.6) <strong>for</strong> the derivative of a product to a function f(x) =u(x)[1/v(x)],we may obtain the derivative of the quotient of two factors. Thus( u) ( ) ′ ′ ( ) )1 1f ′ = = u + u ′ = u(− v′v v v v 2 + u′v ,where (2.12) has been used to evaluate (1/v) ′ . This can now be rearranged intothe more convenient <strong>and</strong> memorisable <strong>for</strong>m( u) ′f ′ vu ′ − uv ′= =v v 2 . (2.13)This can be expressed in words as the derivative of a quotient is equal to the bottomtimes the derivative of the top minus the top times the derivative of the bottom, allover the bottom squared.◮Find the derivative with respect to x of f(x) =sinx/x.Using (2.13) with u(x) =sinx, v(x) =x <strong>and</strong> hence u ′ (x) =cosx, v ′ (x) = 1, we findf ′ (x) =x cos x − sin xx 2= cos xx− sin x . ◭x 22.1.5 Implicit differentiationSo far we have only differentiated functions written in the <strong>for</strong>m y = f(x).However, we may not always be presented with a relationship in this simple<strong>for</strong>m. As an example consider the relation x 3 − 3xy + y 3 = 2. In this case it isnot possible to rearrange the equation to give y as a function of x. Nevertheless,by differentiating term by term with respect to x (implicit differentiation), we canfind the derivative of y.47


PRELIMINARY CALCULUS◮Find dy/dx if x 3 − 3xy + y 3 =2.Differentiating each term in the equation with respect to x we obtainddx (x3 ) − ddx (3xy)+ ddx (y3 )= ddx (2),(⇒ 3x 2 − 3x dy )dx +3y +3y 2 dydx =0,where the derivative of 3xy has been found using the product rule. Hence, rearranging <strong>for</strong>dy/dx,dydx = y − x2y 2 − x .Note that dy/dx is a function of both x <strong>and</strong> y <strong>and</strong> cannot be expressed as a function of xonly. ◭2.1.6 Logarithmic differentiationIn circumstances in which the variable with respect to which we are differentiatingis an exponent, taking logarithms <strong>and</strong> then differentiating implicitly is the simplestway to find the derivative.◮Find the derivative with respect to x of y = a x .To find the required derivative we first take logarithms <strong>and</strong> then differentiate implicitly:ln y =lna x 1 dy= x ln a ⇒y dx =lna.Now, rearranging <strong>and</strong> substituting <strong>for</strong> y, we finddydx = y ln a = ax ln a. ◭2.1.7 Leibnitz’ theoremWe have discussed already how to find the derivative of a product of two or morefunctions. We now consider Leibnitz’ theorem, which gives the correspondingresults <strong>for</strong> the higher derivatives of products.Consider again the function f(x) =u(x)v(x). We know from the product rulethat f ′ = uv ′ + u ′ v. Using the rule once more <strong>for</strong> each of the products, we obtainf ′′ =(uv ′′ + u ′ v ′ )+(u ′ v ′ + u ′′ v)= uv ′′ +2u ′ v ′ + u ′′ v.Similarly, differentiating twice more givesf ′′′ = uv ′′′ +3u ′ v ′′ +3u ′′ v ′ + u ′′′ v,f (4) = uv (4) +4u ′ v ′′′ +6u ′′ v ′′ +4u ′′′ v ′ + u (4) v.48


2.1 DIFFERENTIATIONThe pattern emerging is clear <strong>and</strong> strongly suggests that the results generalise tof (n) =n∑r=0n!r!(n − r)! u(r) v (n−r) =n∑n C r u (r) v (n−r) , (2.14)where the fraction n!/[r!(n − r)!] is identified with the binomial coefficient n C r(see chapter 1). To prove that this is so, we use the method of induction as follows.Assume that (2.14) is valid <strong>for</strong> n equal to some integer N. Thenf (N+1) ===N∑r=0N d (C r u (r) v (N−r))dxr=0N∑N C r [u (r) v (N−r+1) + u (r+1) v (N−r) ]r=0N∑N+1∑N C s u (s) v (N+1−s) +N C s−1 u (s) v (N+1−s) ,s=0where we have substituted summation index s <strong>for</strong> r in the first summation, <strong>and</strong><strong>for</strong> r + 1 in the second. Now, from our earlier discussion of binomial coefficients,equation (1.51), we haves=1N C s + N C s−1 = N+1 C s<strong>and</strong> so, after separating out the first term of the first summation <strong>and</strong> the lastterm of the second, obtainf (N+1) = N C 0 u (0) v (N+1) +N∑N+1 C s u (s) v (N+1−s) + N C N u (N+1) v (0) .s=1But N C 0 =1= N+1 C 0 <strong>and</strong> N C N =1= N+1 C N+1 , <strong>and</strong> so we may writef (N+1) = N+1 C 0 u (0) v (N+1) +N∑N+1 C s u (s) v (N+1−s) + N+1 C N+1 u (N+1) v (0)s=1N+1∑=N+1 C s u (s) v (N+1−s) .s=0This is just (2.14) with n set equal to N + 1. Thus, assuming the validity of (2.14)<strong>for</strong> n = N implies its validity <strong>for</strong> n = N + 1. However, when n =1equation(2.14) is simply the product rule, <strong>and</strong> this we have already proved directly. Theseresults taken together establish the validity of (2.14) <strong>for</strong> all n <strong>and</strong> prove Leibnitz’theorem.49


PRELIMINARY CALCULUSf(x)QABCSFigure 2.2 A graph of a function, f(x), showing how differentiation correspondsto finding the gradient of the function at a particular point. Points B,Q <strong>and</strong> S are stationary points (see text).x◮Find the third derivative of the function f(x) =x 3 sin x.Using (2.14) we immediately findf ′′′ (x) =6sinx +3(6x)cosx +3(3x 2 )(− sin x)+x 3 (− cos x)=3(2− 3x 2 )sinx + x(18 − x 2 )cosx. ◭2.1.8 Special points of a functionWe have interpreted the derivative of a function as the gradient of the function atthe relevant point (figure 2.1). If the gradient is zero <strong>for</strong> some particular value ofx then the function is said to have a stationary point there. Clearly, in graphicalterms, this corresponds to a horizontal tangent to the graph.Stationary points may be divided into three categories <strong>and</strong> an example of eachis shown in figure 2.2. Point B is said to be a minimum since the function increasesin value in both directions away from it. Point Q is said to be a maximum sincethe function decreases in both directions away from it. Note that B is not theoverall minimum value of the function <strong>and</strong> Q is not the overall maximum; rather,they are a local minimum <strong>and</strong> a local maximum. Maxima <strong>and</strong> minima are knowncollectively as turning points.The third type of stationary point is the stationary point of inflection, S. Inthis case the function falls in the positive x-direction <strong>and</strong> rises in the negativex-direction so that S is neither a maximum nor a minimum. Nevertheless, thegradient of the function is zero at S, i.e. the graph of the function is flat there,<strong>and</strong> this justifies our calling it a stationary point. Of course, a point at which the50


2.1 DIFFERENTIATIONgradient of the function is zero but the function rises in the positive x-direction<strong>and</strong> falls in the negative x-direction is also a stationary point of inflection.The above distinction between the three types of stationary point has beenmade rather descriptively. However, it is possible to define <strong>and</strong> distinguish stationarypoints mathematically. From their definition as points of zero gradient,all stationary points must be characterised by df/dx = 0. In the case of theminimum, B, the slope, i.e. df/dx, changes from negative at A to positive at Cthrough zero at B. Thus df/dx is increasing <strong>and</strong> so the second derivative d 2 f/dx 2must be positive. Conversely, at the maximum, Q, we must have that d 2 f/dx 2 isnegative.It is less obvious, but intuitively reasonable, that at S, d 2 f/dx 2 is zero. This maybe inferred from the following observations. To the left of S the curve is concaveupwards so that df/dx is increasing with x <strong>and</strong> hence d 2 f/dx 2 > 0. To the rightof S, however, the curve is concave downwards so that df/dx is decreasing withx <strong>and</strong> hence d 2 f/dx 2 < 0.In summary, at a stationary point df/dx = 0 <strong>and</strong>(i) <strong>for</strong> a minimum, d 2 f/dx 2 > 0,(ii) <strong>for</strong> a maximum, d 2 f/dx 2 < 0,(iii) <strong>for</strong> a stationary point of inflection, d 2 f/dx 2 = 0 <strong>and</strong> d 2 f/dx 2 changes signthrough the point.In case (iii), a stationary point of inflection, in order that d 2 f/dx 2 changes signthrough the point we normally require d 3 f/dx 3 ≠ 0 at that point. This simplerule can fail <strong>for</strong> some functions, however, <strong>and</strong> in general if the first non-vanishingderivative of f(x) at the stationary point is f (n) then if n is even the point is amaximum or minimum <strong>and</strong> if n is odd the point is a stationary point of inflection.This may be seen from the Taylor expansion (see equation (4.17)) of the functionabout the stationary point, but it is not proved here.◮Find the positions <strong>and</strong> natures of the stationary points of the functionf(x) =2x 3 − 3x 2 − 36x +2.The first criterion <strong>for</strong> a stationary point is that df/dx = 0, <strong>and</strong> hence we setdfdx =6x2 − 6x − 36 = 0,from which we obtain(x − 3)(x +2)=0.Hence the stationary points are at x =3<strong>and</strong>x = −2. To determine the nature of thestationary point we must evaluate d 2 f/dx 2 :d 2 f=12x − 6.dx2 51


PRELIMINARY CALCULUSf(x)GxFigure 2.3 The graph of a function f(x) that has a general point of inflectionat the point G.Now, we examine each stationary point in turn. For x =3,d 2 f/dx 2 = 30. Since this ispositive, we conclude that x = 3 is a minimum. Similarly, <strong>for</strong> x = −2, d 2 f/dx 2 = −30 <strong>and</strong>so x = −2 is a maximum. ◭So far we have concentrated on stationary points, which are defined to havedf/dx = 0. We have found that at a stationary point of inflection d 2 f/dx 2 isalso zero <strong>and</strong> changes sign. This naturally leads us to consider points at whichd 2 f/dx 2 is zero <strong>and</strong> changes sign but at which df/dx is not, in general, zero. Suchpoints are called general points of inflection or simply points of inflection. Clearly,a stationary point of inflection is a special case <strong>for</strong> which df/dx is also zero.At a general point of inflection the graph of the function changes from beingconcave upwards to concave downwards (or vice versa), but the tangent to thecurve at this point need not be horizontal. A typical example of a general pointof inflection is shown in figure 2.3.The determination of the stationary points of a function, together with theidentification of its zeros, infinities <strong>and</strong> possible asymptotes, is usually sufficientto enable a graph of the function showing most of its significant features to besketched. Some examples <strong>for</strong> the reader to try are included in the exercises at theend of this chapter.2.1.9 Curvature of a functionIn the previous section we saw that at a point of inflection of the functionf(x), the second derivative d 2 f/dx 2 changes sign <strong>and</strong> passes through zero. Thecorresponding graph of f shows an inversion of its curvature at the point ofinflection. We now develop a more quantitative measure of the curvature of afunction (or its graph), which is applicable at general points <strong>and</strong> not just in theneighbourhood of a point of inflection.As in figure 2.1, let θ be the angle made with the x-axis by the tangent at a52


2.1 DIFFERENTIATIONf(x)C∆θρPQθθ +∆θxFigure 2.4 Two neighbouring tangents to the curve f(x) whose slopes differby ∆θ. The angular separation of the corresponding radii of the circle ofcurvature is also ∆θ.point P on the curve f = f(x), with tan θ = df/dx evaluated at P . Now consideralso the tangent at a neighbouring point Q on the curve, <strong>and</strong> suppose that itmakes an angle θ +∆θ with the x-axis, as illustrated in figure 2.4.It follows that the corresponding normals at P <strong>and</strong> Q, which are perpendicularto the respective tangents, also intersect at an angle ∆θ. Furthermore, their pointof intersection, C in the figure, will be the position of the centre of a circle thatapproximates the arc PQ, at least to the extent of having the same tangents atthe extremities of the arc. This circle is called the circle of curvature.For a finite arc PQ, the lengths of CP <strong>and</strong> CQ will not, in general, be equal,as they would be if f = f(x) were in fact the equation of a circle. But, as Qis allowed to tend to P ,i.e.as∆θ → 0, they do become equal, their commonvalue being ρ, the radius of the circle, known as the radius of curvature. It followsimmediately that the curve <strong>and</strong> the circle of curvature have a common tangentat P <strong>and</strong> lie on the same side of it. The reciprocal of the radius of curvature, ρ −1 ,defines the curvature of the function f(x) at the point P .The radius of curvature can be defined more mathematically as follows. Thelength ∆s of arc PQ is approximately equal to ρ∆θ <strong>and</strong>, in the limit ∆θ → 0, thisrelationship defines ρ as∆sρ = lim∆θ→0 ∆θ = dsdθ . (2.15)It should be noted that, as s increases, θ may increase or decrease according towhether the curve is locally concave upwards (i.e. shaped as if it were near aminimum in f(x)) or concave downwards. This is reflected in the sign of ρ, whichthere<strong>for</strong>e also indicates the position of the curve (<strong>and</strong> of the circle of curvature)53


PRELIMINARY CALCULUSrelative to the common tangent, above or below. Thus a negative value of ρindicates that the curve is locally concave downwards <strong>and</strong> that the tangent liesabove the curve.We next obtain an expression <strong>for</strong> ρ, not in terms of s <strong>and</strong> θ but in termsof x <strong>and</strong> f(x). The expression, though somewhat cumbersome, follows from thedefining equation (2.15), the defining property of θ that tan θ = df/dx ≡ f ′ <strong>and</strong>the fact that the rate of change of arc length with x is given by[ ( ) ] 2 1/2ds dfdx = 1+. (2.16)dxThis last result, simply quoted here, is proved more <strong>for</strong>mally in subsection 2.2.13.From the chain rule (2.11) it follows thatρ = dsdθ = ds dxdx dθ . (2.17)Differentiating both sides of tan θ = df/dx with respect to x givessec 2 θ dθdx = d2 fdx 2 ≡ f′′ ,from which, using sec 2 θ =1+tan 2 θ =1+(f ′ ) 2 , we can obtain dx/dθ asdxdθ = 1+tan2 θf ′′ = 1+(f′ ) 2f ′′ . (2.18)Substituting (2.16) <strong>and</strong> (2.18) into (2.17) then yields the final expression <strong>for</strong> ρ,ρ =[1+(f ′ ) 2] 3/2f ′′ . (2.19)It should be noted that the quantity in brackets is always positive <strong>and</strong> that itsthree-halves root is also taken as positive. The sign of ρ is thus solely determinedby that of d 2 f/dx 2 , in line with our previous discussion relating the sign towhether the curve is concave or convex upwards. If, as happens at a point ofinflection, d 2 f/dx 2 is zero then ρ is <strong>for</strong>mally infinite <strong>and</strong> the curvature of f(x) iszero. As d 2 f/dx 2 changes sign on passing through zero, both the local tangent<strong>and</strong> the circle of curvature change from their initial positions to the opposite sideof the curve.54


2.1 DIFFERENTIATION◮Show that the radius of curvature at the point (x, y) on the ellipsex 2a + y22 b =1 2has magnitude (a 4 y 2 + b 4 x 2 ) 3/2 /(a 4 b 4 ) <strong>and</strong> the opposite sign to y. Check the special caseb = a, <strong>for</strong> which the ellipse becomes a circle.Differentiating the equation of the ellipse with respect to x gives2xa + 2y dy2 b 2 dx =0<strong>and</strong> sodydx = − b2 xa 2 y .A second differentiation, using (2.13), then yields( ) ( )d 2 ydx = − b2 y − xy′= − b4 y22 a 2 y 2 a 2 y 3 b + x2= − b42 a 2 a 2 y , 3where we have used the fact that (x, y) lies on the ellipse. We note that d 2 y/dx 2 , <strong>and</strong> henceρ, has the opposite sign to y 3 <strong>and</strong> hence to y. Substituting in (2.19) gives <strong>for</strong> the magnitudeof the radius of curvature[ 1+b 4 x 2 /(a 4 y 2 ) ] 3/2|ρ| =∣ −b 4 /(a 2 y 3 ) ∣ = (a4 y 2 + b 4 x 2 ) 3/2.a 4 b 4For the special case b = a, |ρ| reduces to a −2 (y 2 + x 2 ) 3/2 <strong>and</strong>, since x 2 + y 2 = a 2 ,thisinturn gives |ρ| = a, as expected. ◭The discussion in this section has been confined to the behaviour of curvesthat lie in one plane; examples of the application of curvature to the bending ofloaded beams <strong>and</strong> to particle orbits under the influence of a central <strong>for</strong>ces can befound in the exercises at the ends of later chapters. A more general treatment ofcurvature in three dimensions is given in section 10.3, where a vector approach isadopted.2.1.10 Theorems of differentiationRolle’s theoremRolle’s theorem (figure 2.5) states that if a function f(x) is continuous in therange a ≤ x ≤ c, is differentiable in the range a


PRELIMINARY CALCULUSf(x)a b cxFigure 2.5 The graph of a function f(x), showing that if f(a) =f(c) thenatone point at least between x = a <strong>and</strong> x = c the graph has zero gradient.f(x)f(c)Cf(a)Aa b cxFigure 2.6 The graph of a function f(x); at some point x = b it has the samegradient as the line AC.Mean value theoremThe mean value theorem (figure 2.6) states that if a function f(x) is continuousin the range a ≤ x ≤ c <strong>and</strong> differentiable in the range a


2.1 DIFFERENTIATION<strong>and</strong> hence the difference between the curve <strong>and</strong> the line ish(x) =f(x) − g(x) =f(x) − f(a) − (x − a)f(c) − f(a).c − aSince the curve <strong>and</strong> the line intersect at A <strong>and</strong> C, h(x) = 0 at both of these points.Hence, by an application of Rolle’s theorem, h ′ (x) = 0 <strong>for</strong> at least one point bbetween A <strong>and</strong> C. Differentiating our expression <strong>for</strong> h(x), we find<strong>and</strong> hence at b, whereh ′ (x) =0,h ′ (x) =f ′ (x) −f ′ (b) =f(c) − f(a),c − af(c) − f(a).c − aApplications of Rolle’s theorem <strong>and</strong> the mean value theoremSince the validity of Rolle’s theorem is intuitively obvious, given the conditionsimposed on f(x), it will not be surprising that the problems that can be solvedby applications of the theorem alone are relatively simple ones. Nevertheless wewill illustrate it with the following example.◮What semi-quantitative results can be deduced by applying Rolle’s theorem to the followingfunctions f(x), witha <strong>and</strong> c chosen so that f(a) =f(c) =0?(i)sinx, (ii) cos x,(iii)x 2 − 3x +2, (iv) x 2 +7x +3, (v) 2x 3 − 9x 2 − 24x + k.(i) If the consecutive values of x that make sin x =0areα 1 ,α 2 ,... (actually x = nπ, <strong>for</strong>any integer n) then Rolle’s theorem implies that the derivative of sin x, namelycosx, hasat least one zero lying between each pair of values α i <strong>and</strong> α i+1 .(ii) In an exactly similar way, we conclude that the derivative of cos x, namely− sin x,has at least one zero lying between consecutive pairs of zeros of cos x. Thesetworesultstaken together (but neither separately) imply that sin x <strong>and</strong> cos x have interleavingzeros.(iii) For f(x) =x 2 − 3x +2, f(a) =f(c) =0ifa <strong>and</strong> c are taken as 1 <strong>and</strong> 2 respectively.Rolle’s theorem then implies that f ′ (x) =2x − 3=0hasasolutionx = b with b in therange 1


PRELIMINARY CALCULUSIn each case, as might be expected, the application of Rolle’s theorem does no more thanfocus attention on particular ranges of values; it does not yield precise answers. ◭Direct verification of the mean value theorem is straight<strong>for</strong>ward when it isapplied to simple functions. For example, if f(x) =x 2 , it states that there is avalue b in the interval a


2.2 INTEGRATIONf(x)abxFigure 2.7An integral as the area under a curve.2.2 IntegrationThe notion of an integral as the area under a curve will be familiar to the reader.In figure 2.7, in which the solid line is a plot of a function f(x), the shaded arearepresents the quantity denoted byI =∫ baf(x) dx. (2.21)This expression is known as the definite integral of f(x) betweenthelower limitx = a <strong>and</strong> the upper limit x = b, <strong>and</strong>f(x) is called the integr<strong>and</strong>.2.2.1 Integration from first principlesThe definition of an integral as the area under a curve is not a <strong>for</strong>mal definition,but one that can be readily visualised. The <strong>for</strong>mal definition of I involvessubdividing the finite interval a ≤ x ≤ b into a large number of subintervals, bydefining intermediate points ξ i such that a = ξ 0


PRELIMINARY CALCULUSf(x)ax 1 x 2 x 3 x 4 x 5ξ 1 ξ 2 ξ 3 ξ 4bxFigure 2.8 The evaluation of a definite integral by subdividing the intervala ≤ x ≤ b into subintervals.◮Evaluate from first principles the integral I = ∫ b0 x2 dx.We first approximate the area under the curve y = x 2 between 0 <strong>and</strong> b by n rectangles ofequal width h. If we take the value at the lower end of each subinterval (in the limit of aninfinite number of subintervals we could equally well have chosen the value at the upperend) to give the height of the corresponding rectangle, then the area of the kth rectanglewill be (kh) 2 h = k 2 h 3 . The total area is thus∑n−1A = k 2 h 3 =(h 3 ) 1 n(n − 1)(2n − 1),6k=0where we have used the expression <strong>for</strong> the sum of the squares of the natural numbersderived in subsection 1.7.1. Now h = b/n <strong>and</strong> so( ) (b3 n b3A = (n − 1)(2n − 1) = 1 − 1 )(2 − 1 ).n 3 6 6 n nAs n →∞, A → b 3 /3, which is thus the value I of the integral. ◭Some straight<strong>for</strong>ward properties of definite integrals that are almost self-evidentare as follows:∫ ba∫ ca∫ ba0 dx =0,f(x) dx =∫ b[f(x)+g(x)] dx =a∫ aaf(x) dx +∫ ba60f(x) dx =0, (2.23)∫ cbf(x) dx +f(x) dx, (2.24)∫ bag(x) dx. (2.25)


2.2 INTEGRATIONCombining (2.23) <strong>and</strong> (2.24) with c set equal to a shows that∫ bf(x) dx = −∫ aabf(x) dx. (2.26)2.2.2 Integration as the inverse of differentiationThe definite integral has been defined as the area under a curve between twofixed limits. Let us now consider the integralF(x) =∫ xaf(u) du (2.27)in which the lower limit a remains fixed but the upper limit x is now variable. Itwill be noticed that this is essentially a restatement of (2.21), but that the variablex in the integr<strong>and</strong> has been replaced by a new variable u. It is conventional torename the dummy variable in the integr<strong>and</strong> in this way in order that the samevariable does not appear in both the integr<strong>and</strong> <strong>and</strong> the integration limits.It is apparent from (2.27) that F(x) is a continuous function of x, but at firstglance the definition of an integral as the area under a curve does not connect withour assertion that integration is the inverse process to differentiation. However,by considering the integral (2.27) <strong>and</strong> using the elementary property (2.24), weobtainF(x +∆x) ==∫ x+∆xa∫ xa= F(x)+f(u) duf(u) du +x∫ x+∆xRearranging <strong>and</strong> dividing through by ∆x yieldsF(x +∆x) − F(x)∆xx= 1∆x∫ x+∆xf(u) du.∫ x+∆xxf(u) duf(u) du.Letting ∆x → 0 <strong>and</strong> using (2.1) we find that the LHS becomes dF/dx, whereasthe RHS becomes f(x). The latter conclusion follows because when ∆x is smallthe value of the integral on the RHS is approximately f(x)∆x, <strong>and</strong> in the limit∆x → 0 no approximation is involved. ThusdF(x)= f(x), (2.28)dxor, substituting <strong>for</strong> F(x) from (2.27),[∫d x]f(u) du = f(x).dxa61


PRELIMINARY CALCULUSFrom the last two equations it is clear that integration can be considered asthe inverse of differentiation. However, we see from the above analysis that thelower limit a is arbitrary <strong>and</strong> so differentiation does not have a unique inverse.Any function F(x) obeying (2.28) is called an indefinite integral of f(x), thoughany two such functions can differ by at most an arbitrary additive constant. Sincethe lower limit is arbitrary, it is usual to writeF(x) =∫ xf(u) du (2.29)<strong>and</strong> explicitly include the arbitrary constant only when evaluating F(x). Theevaluation is conventionally written in the <strong>for</strong>m∫f(x) dx = F(x)+c (2.30)where c is called the constant of integration. It will be noticed that, in the absenceof any integration limits, we use the same symbol <strong>for</strong> the arguments of both f<strong>and</strong> F. This can be confusing, but is sufficiently common practice that the readerneeds to become familiar with it.We also note that the definite integral of f(x) between the fixed limits x = a<strong>and</strong> x = b can be written in terms of F(x). From (2.27) we have∫ baf(x) dx =∫ bx 0f(x) dx −∫ ax 0f(x) dx= F(b) − F(a), (2.31)where x 0 is any third fixed point. Using the notation F ′ (x) =dF/dx, wemayrewrite (2.28) as F ′ (x) =f(x), <strong>and</strong> so express (2.31) as∫ baF ′ (x) dx = F(b) − F(a) ≡ [F] b a.In contrast to differentiation, where repeated applications of the product rule<strong>and</strong>/or the chain rule will always give the required derivative, it is not alwayspossible to find the integral of an arbitrary function. Indeed, in most real physicalproblems exact integration cannot be per<strong>for</strong>med <strong>and</strong> we have to revert tonumerical approximations. Despite this cautionary note, it is in fact possible tointegrate many simple functions <strong>and</strong> the following subsections introduce the mostcommon types. Many of the techniques will be familiar to the reader <strong>and</strong> so aresummarised by example.2.2.3 Integration by inspectionThe simplest method of integrating a function is by inspection. Some of the moreelementary functions have well-known integrals that should be remembered. Thereader will notice that these integrals are precisely the inverses of the derivatives62


2.2 INTEGRATIONfound near the end of subsection 2.1.1. A few are presented below, using the <strong>for</strong>mgiven in (2.30):∫∫adx= ax + c, ax n dx = axn+1n +1 + c,∫∫e ax dx = eaxaa + c, dx = a ln x + c,x∫∫a sin bx−a cos bxa cos bx dx = + c, a sin bx dx = + c,bb∫∫a tan bx dx =−a ln(cos bx)b+ c,a(a 2 + x 2 dx x)=tan−1 + c,a∫−1( √a2 − x dx x)2=cos−1 + c,a∫∫a cos bx sin n bx dx = a sinn+1 bxb(n +1)a sin bx cos n bx dx = −a cosn+1 bxb(n +1)∫+ c,+ c,1( √a2 − x dx x)2=sin−1 + c,awhere the integrals that depend on n are valid <strong>for</strong> all n ≠ −1 <strong>and</strong>wherea <strong>and</strong> bare constants. In the two final results |x| ≤a.2.2.4 Integration of sinusoidal functionsIntegrals of the type ∫ sin n xdx <strong>and</strong> ∫ cos n xdx may be found by using trigonometricexpansions. Two methods are applicable, one <strong>for</strong> odd n <strong>and</strong> the other <strong>for</strong>even n. They are best illustrated by example.◮Evaluate the integral I = ∫ sin 5 xdx.Rewriting the integral as a product of sin x <strong>and</strong> an even power of sin x, <strong>and</strong> then usingthe relation sin 2 x =1− cos 2 x yields∫I = sin 4 x sin xdx∫= (1 − cos 2 x) 2 sin xdx∫= (1 − 2cos 2 x +cos 4 x)sinxdx∫= (sin x − 2sinx cos 2 x +sinx cos 4 x) dx= − cos x + 2 3 cos3 x − 1 5 cos5 x + c,where the integration has been carried out using the results of subsection 2.2.3. ◭63


PRELIMINARY CALCULUS◮Evaluate the integral I = ∫ cos 4 xdx.Rewriting the integral as a power of cos 2 x <strong>and</strong> then using the double-angle <strong>for</strong>mulacos 2 x = 1 (1+cos2x) yields2∫∫ ( ) 2 1+cos2xI = (cos 2 x) 2 dx =dx2∫1= (1+2cos2x 4 +cos2 2x) dx.Using the double-angle <strong>for</strong>mula again we may write cos 2 2x = 1 (1+cos4x), <strong>and</strong> hence2∫ [I = 1 + 1 cos 2x + 1 (1+cos4x)] dx4 2 8= 1 x + 1 sin 2x + 1 x + 1 sin 4x + c4 4 8 32= 3 x + 1 1sin 2x + sin 4x + c. ◭8 4 322.2.5 Logarithmic integrationIntegrals <strong>for</strong> which the integr<strong>and</strong> may be written as a fraction in which thenumerator is the derivative of the denominator may be evaluated using∫ f ′ (x)dx =lnf(x)+c. (2.32)f(x)This follows directly from the differentiation of a logarithm as a function of afunction (see subsection 2.1.3).◮Evaluate the integral∫I =6x 2 +2cosxx 3 +sinxdx.We note first that the numerator can be factorised to give 2(3x 2 +cosx), <strong>and</strong> then thatthe quantity in brackets is the derivative of the denominator. Hence∫ 3x 2 +cosxI =2x 3 +sinx dx =2ln(x3 +sinx)+c. ◭2.2.6 Integration using partial fractionsThe method of partial fractions was discussed at some length in section 1.4, butin essence consists of the manipulation of a fraction (here the integr<strong>and</strong>) in sucha way that it can be written as the sum of two or more simpler fractions. Againwe illustrate the method by an example.64


2.2 INTEGRATION◮Evaluate the integral∫I =1x 2 + x dx.We note that the denominator factorises to give x(x + 1). Hence∫1I =x(x +1) dx.We now separate the fraction into two partial fractions <strong>and</strong> integrate directly:∫ ( 1I =x − 1 )( ) xdx =lnx − ln(x +1)+c =ln + c. ◭x +1x +12.2.7 Integration by substitutionSometimes it is possible to make a substitution of variables that turns a complicatedintegral into a simpler one, which can then be integrated by a st<strong>and</strong>ardmethod. There are many useful substitutions <strong>and</strong> knowing which to use is a matterof experience. We now present a few examples of particularly useful substitutions.◮Evaluate the integral∫I =1√1 − x2 dx.Making the substitution x =sinu, we note that dx =cosudu, <strong>and</strong> hence∫∫∫1I = √1 − sin 2 u cos udu= 1√cos2 u cos udu= du = u + c.Now substituting back <strong>for</strong> u,I =sin −1 x + c.This corresponds to one of the results given in subsection 2.2.3. ◭Another particular example of integration by substitution is af<strong>for</strong>ded by integralsof the <strong>for</strong>m∫∫1I =a + b cos x dx or I = 1dx. (2.33)a + b sin xIn these cases, making the substitution t = tan(x/2) yields integrals that canbe solved more easily than the originals. Formulae expressing sin x <strong>and</strong> cos x interms of t were derived in equations (1.32) <strong>and</strong> (1.33) (see p. 14), but be<strong>for</strong>e wecan use them we must relate dx to dt as follows.65


PRELIMINARY CALCULUSSincethe required relationship isdtdx = 1 x 2 sec2 2 = 1 (1+tan 2 x )= 1+t2 ,2 2 2dx = 2 dt. (2.34)1+t2 ◮Evaluate the integral∫I =21+3cosx dx.Rewriting cos x in terms of t <strong>and</strong> using (2.34) yields∫I =∫=∫=∫=21+3 [ (1 − t 2 )(1 + t 2 ) −1] ( 21+t 2 )dt2(1 + t 2 )1+t 2 +3(1− t 2 )∫22 − t dt = 2( 21+t 2 )dt2( √ 2 − t)( √ 2+t) dt(1 1√ √ + 1 )√ dt2 2 − t 2+t= − 1 √2ln( √ 2 − t)+1 √2ln( √ 2+t)+c[= √ 1√ ]2+tan(x/2)ln √ + c. ◭2 2 − tan (x/2)Integrals of a similar <strong>for</strong>m to (2.33), but involving sin 2x, cos2x, tan 2x, sin 2 x,cos 2 x or tan 2 x instead of cos x <strong>and</strong> sin x, should be evaluated by using thesubstitution t =tanx. Inthiscasesin x =t√1+t2 , cos x = 1√1+t2<strong>and</strong> dx = dt1+t 2 . (2.35)A final example of the evaluation of integrals using substitution is the methodof completing the square (cf. subsection 1.7.3).66


2.2 INTEGRATION◮Evaluate the integral∫I =1x 2 +4x +7 dx.We can write the integral in the <strong>for</strong>m∫I =1(x +2) 2 +3 dx.Substituting y = x + 2, we find dy = dx <strong>and</strong> hence∫1I =y 2 +3 dy,Hence, by comparison with the table of st<strong>and</strong>ard integrals (see subsection 2.2.3)√ ( ) √ ( ) 3 y√3 3 x +2I =3 tan−1 + c =3 tan−1 √ + c. ◭32.2.8 Integration by partsIntegration by parts is the integration analogy of product differentiation. Theprinciple is to break down a complicated function into two functions, at least oneof which can be integrated by inspection. The method in fact relies on the result<strong>for</strong> the differentiation of a product. Recalling from (2.6) thatd dv(uv) =udx dx + dudx v,where u <strong>and</strong> v are functions of x, we now integrate to find∫uv = u dv ∫ dudx dx + dx vdx.Rearranging into the st<strong>and</strong>ard <strong>for</strong>m <strong>for</strong> integration by parts gives∫u dv ∫ dudx dx = uv − vdx. (2.36)dxIntegration by parts is often remembered <strong>for</strong> practical purposes in the <strong>for</strong>mthe integral of a product of two functions is equal to {the first times the integral ofthe second} minus the integral of {the derivative of the first times the integral ofthe second}. Here, u is ‘the first’ <strong>and</strong> dv/dx is ‘the second’; clearly the integral vof ‘the second’ must be determinable by inspection.◮Evaluate the integral I = ∫ x sin x dx.In the notation given above, we identify x with u <strong>and</strong> sin x with dv/dx. Hence v = − cos x<strong>and</strong> du/dx = 1 <strong>and</strong> so using (2.36)∫I = x(− cos x) − (1)(− cos x) dx = −x cos x +sinx + c. ◭67


PRELIMINARY CALCULUSThe separation of the functions is not always so apparent, as is illustrated bythe following example.◮Evaluate the integral I = ∫ x 3 e −x2 dx.Firstly we rewrite the integral as∫I =x 2 (xe −x2) dx.Now, using the notation given above, we identify x 2 with u <strong>and</strong> xe −x2 with dv/dx. Hencev = − 1 2 e−x2 <strong>and</strong> du/dx =2x, sothat∫I = − 1 2 x2 e −x2 − (−x)e −x2 dx = − 1 2 x2 e −x2 − 1 2 e−x2 + c. ◭A trick that is sometimes useful is to take ‘1’ as one factor of the product, asis illustrated by the following example.◮Evaluate the integral I = ∫ ln xdx.Firstly we rewrite the integral as∫I = (ln x)1dx.Now, using the notation above, we identify ln x with u <strong>and</strong> 1 with dv/dx. Hence we havev = x <strong>and</strong> du/dx =1/x, <strong>and</strong>so∫ ( ) 1I =(lnx)(x) − xdx= x ln x − x + c. ◭xIt is sometimes necessary to integrate by parts more than once. In doing so,we may occasionally re-encounter the original integral I. In such cases we canobtain a linear algebraic equation <strong>for</strong> I that can be solved to obtain its value.◮Evaluate the integral I = ∫ e ax cos bx dx.Integrating by parts, taking e ax as the first function, we find( ) ∫ ( )sin bxsin bxI = e ax − ae ax dx,bbwhere, <strong>for</strong> convenience, we have omitted the constant of integration. Integrating by partsa second time,( ) ( ) ∫ ( )sin bx − cos bx− cos bxI = e ax − ae ax + a 2 e ax dx.bb 2 b 2Notice that the integral on the RHS is just −a 2 /b 2 times the original integral I. Thus( 1I = e ax b sin bx + a )b cos bx − a22 b I. 268


2.2 INTEGRATIONRearranging this expression to obtain I explicitly <strong>and</strong> including the constant of integrationwe findI =eax(b sin bx + a cos bx)+c. (2.37)a 2 + b2 Another method of evaluating this integral, using the exponential of a complex number,is given in section 3.6. ◭2.2.9 Reduction <strong>for</strong>mulaeIntegration using reduction <strong>for</strong>mulae is a process that involves first evaluating asimple integral <strong>and</strong> then, in stages, using it to find a more complicated integral.◮Using integration by parts, find a relationship between I n <strong>and</strong> I n−1 whereI n =∫ 10(1 − x 3 ) n dx<strong>and</strong> n is any positive integer. Hence evaluate I 2 = ∫ 10 (1 − x3 ) 2 dx.Writing the integr<strong>and</strong> as a product <strong>and</strong> separating the integral into two we findI n ==∫ 10∫ 10(1 − x 3 )(1 − x 3 ) n−1 dx(1 − x 3 ) n−1 dx −∫ 10x 3 (1 − x 3 ) n−1 dx.The first term on the RHS is clearly I n−1 <strong>and</strong> so, writing the integr<strong>and</strong> in the second termon the RHS as a product,Integrating by parts we findI n = I n−1 +I n = I n−1 −∫ 10(x)x 2 (1 − x 3 ) n−1 dx.[ x] 1∫ 13n (1 − x3 ) n −0 013n (1 − x3 ) n dx= I n−1 +0− 13n I n,which on rearranging givesI n =3n3n +1 I n−1.We now have a relation connecting successive integrals. Hence, if we can evaluate I 0 ,wecan find I 1 , I 2 etc. Evaluating I 0 is trivial:I 0 =∫ 10(1 − x 3 ) 0 dx =∫ 10dx = [x] 1 0 =1.Hence(3 × 1)I 1 =(3 × 1) + 1 × 1= 3 4 , I (3 × 2)2 =(3 × 2) + 1 × 3 4 = 9 14 .Although the first few I n could be evaluated by direct multiplication, this becomes tedious<strong>for</strong> integrals containing higher values of n; these are there<strong>for</strong>e best evaluated using thereduction <strong>for</strong>mula. ◭69


PRELIMINARY CALCULUS2.2.10 Infinite <strong>and</strong> improper integralsThe definition of an integral given previously does not allow <strong>for</strong> cases in whicheither of the limits of integration is infinite (an infinite integral) or <strong>for</strong> casesin which f(x) is infinite in some part of the range (an improper integral), e.g.f(x) = (2 − x) −1/4 near the point x = 2. Nevertheless, modification of thedefinition of an integral gives infinite <strong>and</strong> improper integrals each a meaning.In the case of an integral I = ∫ baf(x) dx, the infinite integral, in which b tendsto ∞, is defined byI =∫ ∞af(x) dx = lim∫ bb→∞ af(x) dx = limb→∞F(b) − F(a).As previously, F(x) is the indefinite integral of f(x) <strong>and</strong> lim b→∞ F(b) meansthelimit (or value) that F(b) approaches as b →∞; it is evaluated after calculatingthe integral. The <strong>for</strong>mal concept of a limit will be introduced in chapter 4.◮Evaluate the integral∫ ∞xI =0 (x 2 + a 2 ) dx. 2Integrating, we find F(x) =− 1 2 (x2 + a 2 ) −1 + c <strong>and</strong> so]I = lim−b→∞[−12(b 2 + a 2 )( −12a 2 )= 12a 2 . ◭For the case of improper integrals, we adopt the approach of excluding theunbounded range from the integral. For example, if the integr<strong>and</strong> f(x) is infiniteat x = c (say), a ≤ c ≤ b then∫ b∫ c−δ∫ bf(x) dx = lim f(x) dx + lim f(x) dx.aδ→0 aɛ→0 c+ɛ◮Evaluate the integral I = ∫ 20 (2 − x)−1/4 dx.Integrating directly,[I = lim −4(2 − x)3/4] 2−ɛ [= limɛ→03 0 −4ɛ3/4] + 4ɛ→03 3 23/4 = ( )43 2 3/4 . ◭2.2.11 Integration in plane polar coordinatesIn plane polar coordinates ρ, φ, a curve is defined by its distance ρ from theorigin as a function of the angle φ between the line joining a point on the curveto the origin <strong>and</strong> the x-axis, i.e. ρ = ρ(φ). The area of an element is given by70


2.2 INTEGRATIONyCρ(φ + dφ)dAρ(φ)ρdφBOxFigure 2.9 Finding the area of a sector OBC defined by the curve ρ(φ) <strong>and</strong>the radii OB, OC, at angles to the x-axis φ 1 , φ 2 respectively.dA = 1 2 ρ2 dφ, as illustrated in figure 2.9, <strong>and</strong> hence the total area between twoangles φ 1 <strong>and</strong> φ 2 is given by∫ φ21A =2 ρ2 dφ. (2.38)φ 1An immediate observation is that the area of a circle of radius a is given byA =∫ 2π012 a2 dφ = [ 12 a2 φ ] 2π= πa 2 .0◮The equation in polar coordinates of an ellipse with semi-axes a <strong>and</strong> b is1ρ = cos2 φ+ sin2 φ.2 a 2 b 2Find the area A of the ellipse.Using (2.38) <strong>and</strong> symmetry, we have∫ 2πA = 1 a 2 b 2∫ π/22 0 b 2 cos 2 φ + a 2 sin 2 φ dφ =2a2 b 2 0To evaluate this integral we write t =tanφ <strong>and</strong> use (2.35):∫ ∞A =2a 2 b 20∫1∞b 2 + a 2 t dt 2 =2b201b 2 cos 2 φ + a 2 sin 2 φ dφ.1(b/a) 2 + t 2 dt.Finally, from the list of st<strong>and</strong>ard integrals (see subsection 2.2.3),A =2b 2 [ 1(b/a) tan−1] ∞t=2ab(b/a)0( π2 − 0 )= πab. ◭71


PRELIMINARY CALCULUS2.2.12 Integral inequalitiesConsider the functions f(x), φ 1 (x) <strong>and</strong>φ 2 (x) such that φ 1 (x) ≤ f(x) ≤ φ 2 (x) <strong>for</strong>all x in the range a ≤ x ≤ b. It immediately follows that∫ baφ 1 (x) dx ≤∫ baf(x) dx ≤∫ baφ 2 (x) dx, (2.39)which gives us a way of estimating an integral that is difficult to evaluate explicitly.◮Show that the value of the integrallies between 0.810 <strong>and</strong> 0.882.I =∫ 101dx(1 + x 2 + x 3 )1/2We note that <strong>for</strong> x in the range 0 ≤ x ≤ 1, 0 ≤ x 3 ≤ x 2 .Hence(1 + x 2 ) 1/2 ≤ (1 + x 2 + x 3 ) 1/2 ≤ (1+2x 2 ) 1/2 ,<strong>and</strong> so1(1 + x 2 ) ≥ 11/2 (1 + x 2 + x 3 ) ≥ 11/2 (1+2x 2 ) . 1/2Consequently,∫ 10∫11(1 + x 2 ) dx ≥ 1/2from which we obtain[ln(x + √ ] 11+x 2 ) ≥ I ≥00∫11(1 + x 2 + x 3 ) dx ≥ 1/2[√12ln0.8814 ≥ I ≥ 0.81050.882 ≥ I ≥ 0.810.01dx,(1+2x 2 )1/2( √ )] 11x + + 2 x2 0In the last line the calculated values have been rounded to three significant figures,one rounded up <strong>and</strong> the other rounded down so that the proved inequality cannot beunknowingly made invalid. ◭2.2.13 Applications of integrationMean value of a functionThe mean value m of a function between two limits a <strong>and</strong> b is defined bym = 1 ∫ bf(x) dx. (2.40)b − a aThe mean value may be thought of as the height of the rectangle that has thesame area (over the same interval) as the area under the curve f(x). This isillustrated in figure 2.10.72


2.2 INTEGRATIONf(x)mabxFigure 2.10The mean value m of a function.◮Find the mean value m of the function f(x) =x 2 between the limits x =2<strong>and</strong> x =4.Using (2.40),m = 14 − 2∫ 42x 2 dx = 1 [ x32 3] 42= 1 2( 433 − 233)= 283 . ◭Finding the length of a curveFinding the area between a curve <strong>and</strong> certain straight lines provides one exampleof the use of integration. Another is in finding the length of a curve. If a curveis defined by y = f(x) then the distance along the curve, ∆s, that corresponds tosmall changes ∆x <strong>and</strong> ∆y in x <strong>and</strong> y is given by∆s ≈ √ (∆x) 2 +(∆y) 2 ; (2.41)this follows directly from Pythagoras’ theorem (see figure 2.11). Dividing (2.41)through by ∆x <strong>and</strong> letting ∆x → 0weobtain §√( )dsdy 2dx = 1+ .dxClearly the total length s of the curve between the points x = a <strong>and</strong> x = b is thengiven by integrating both sides of the equation:s =∫ ba√1+( ) 2 dydx. (2.42)dx§ Instead of considering small changes ∆x <strong>and</strong> ∆y <strong>and</strong> letting these tend to zero, we could havederived (2.41) by considering infinitesimal changes dx <strong>and</strong> dy from the start. After writing (ds) 2 =(dx) 2 +(dy) 2 , (2.41) may be deduced by using the <strong>for</strong>mal device of dividing through by dx. Althoughnot mathematically rigorous, this method is often used <strong>and</strong> generally leads to the correct result.73


PRELIMINARY CALCULUSf(x)y = f(x)∆s∆x∆yxFigure 2.11 The distance moved along a curve, ∆s, corresponding to thesmall changes ∆x <strong>and</strong> ∆y.In plane polar coordinates,ds = √ (dr) 2 +(rdφ) 2 ⇒ s =∫ r2r 1√1+r 2 ( dφdr) 2dr.(2.43)◮Find the length of the curve y = x 3/2 from x =0to x =2.Using (2.42) <strong>and</strong> noting that dy/dx = 3 2√ x, the length s of the curve is given by∫ 2s =[=023= 827√1+ 9 4 xdx( 4)(9 1+9[ ( 112] x) 3/2240) ] 3/2− 1 . ◭[ (1+ ]= 8 9x) 3/2227 40Surfaces of revolutionConsider the surface S <strong>for</strong>med by rotating the curve y = f(x) about the x-axis(see figure 2.12). The surface area of the ‘collar’ <strong>for</strong>med by rotating an elementof the curve, ds, about the x-axis is 2πy ds, <strong>and</strong> hence the total surface area isS =∫ ba2πy ds.Since (ds) 2 =(dx) 2 +(dy) 2 from (2.41), the total surface area between the planesx = a <strong>and</strong> x = b is√∫ b( ) 2 dyS = 2πy 1+ dx. (2.44)dxa74


2.2 INTEGRATIONydsf(x)adxVbxSFigure 2.12The surface <strong>and</strong> volume of revolution <strong>for</strong> the curve y = f(x).◮Find the surface area of a cone <strong>for</strong>med by rotating about the x-axis the line y =2xbetween x =0<strong>and</strong> x = h.Using (2.44), the surface area is given byS ===∫ h0∫ h0(2π)2x√1+[ ddx (2x) ] 2dx4πx ( 1+2 2) ∫ h1/2dx = 4 √ 5πx dx[2 √ 5πx 2 ] h0=2 √ 5π(h 2 − 0) = 2 √ 5πh 2 . ◭0We note that a surface of revolution may also be <strong>for</strong>med by rotating a lineabout the y-axis. In this case the surface area between y = a <strong>and</strong> y = b isS =∫ ba√( ) dx 22πx 1+ dy. (2.45)dyVolumes of revolutionThe volume V enclosed by rotating the curve y = f(x) about the x-axis can alsobe found (see figure 2.12). The volume of the disc between x <strong>and</strong> x + dx is givenby dV = πy 2 dx. Hence the total volume between x = a <strong>and</strong> x = b isV =∫ ba75πy 2 dx. (2.46)


PRELIMINARY CALCULUS◮Find the volume of a cone enclosed by the surface <strong>for</strong>med by rotating about the x-axisthe line y =2x between x =0<strong>and</strong> x = h.Using (2.46), the volume is given byV =∫ h0π(2x) 2 dx =∫ h04πx 2 dx= [ 43 πx3] h0 = 4 3 π(h3 − 0) = 4 3 πh3 . ◭As be<strong>for</strong>e, it is also possible to <strong>for</strong>m a volume of revolution by rotating a curveabout the y-axis. In this case the volume enclosed between y = a <strong>and</strong> y = b isV =∫ baπx 2 dy. (2.47)2.3 Exercises2.1 Obtain the following derivatives from first principles:(a) the first derivative of 3x +4;(b) the first, second <strong>and</strong> third derivatives of x 2 + x;(c) the first derivative of sin x.2.2 Find from first principles the first derivative of (x+3) 2 <strong>and</strong> compare your answerwith that obtained using the chain rule.2.3 Find the first derivatives of(a) x 2 exp x, (b)2sinx cos x, (c)sin2x, (d)x sin ax,(e) (exp ax)(sin ax)tan −1 ax, (f)ln(x a + x −a ),(g) ln(a x + a −x ), (h) x x .2.4 Find the first derivatives of(a) x/(a + x) 2 ,(b)x/(1 − x) 1/2 ,(c)tanx, assinx/ cos x,(d) (3x 2 +2x +1)/(8x 2 − 4x +2).2.5 Use result (2.12) to find the first derivatives of(a) (2x +3) −3 ,(b)sec 2 x,(c)cosech 3 3x, (d)1/ ln x, (e)1/[sin −1 (x/a)].2.6 Show that the function y(x) =exp(−|x|) defined by⎧⎪⎨ exp x <strong>for</strong> x0,is not differentiable at x = 0. Consider the limiting process <strong>for</strong> both ∆x >0<strong>and</strong>∆x


2.3 EXERCISES2.10 The function y(x) is defined by y(x) =(1+x m ) n .(a) Use the chain rule to show that the first derivative of y is nmx m−1 (1 + x m ) n−1 .(b) The binomial expansion (see section 1.5) of (1 + z) n is(1 + z) n =1+nz +n(n − 1)z 2 + ···+2!n(n − 1) ···(n − r +1)z r + ··· .r!Keeping only the terms of zeroth <strong>and</strong> first order in dx, apply this result twiceto derive result (a) from first principles.(c) Exp<strong>and</strong> y in a series of powers of x be<strong>for</strong>e differentiating term by term.Show that the result is the series obtained by exp<strong>and</strong>ing the answer given<strong>for</strong> dy/dx in (a).2.11 Show by differentiation <strong>and</strong> substitution that the differential equation4x 2 d2 y dy− 4xdx2 dx +(4x2 +3)y =0has a solution of the <strong>for</strong>m y(x) =x n sin x, <strong>and</strong> find the value of n.2.12 Find the positions <strong>and</strong> natures of the stationary points of the following functions:(a) x 3 − 3x +3;(b)x 3 − 3x 2 +3x; (c)x 3 +3x +3;(d) sin ax with a ≠0;(e)x 5 + x 3 ;(f)x 5 − x 3 .2.13 Show that the lowest value taken by the function 3x 4 +4x 3 − 12x 2 +6is−26.2.14 By finding their stationary points <strong>and</strong> examining their general <strong>for</strong>ms, determinethe range of values that each of the following functions y(x) can take. In eachcase make a sketch-graph incorporating the features you have identified.(a) y(x) =(x − 1)/(x 2 +2x +6).(b) y(x) =1/(4+3x − x 2 ).(c) y(x) =(8sinx)/(15 + 8 tan 2 x).2.15 Show that y(x) =xa 2x exp x 2 has no stationary points other than x =0,ifexp(− √ 2)


PRELIMINARY CALCULUSOcCρr +∆rrρQpp +∆pPFigure 2.13 The coordinate system described in exercise 2.20.2.20 A two-dimensional coordinate system useful <strong>for</strong> orbit problems is the tangentialpolarcoordinate system (figure 2.13). In this system a curve is defined by r, thedistance from a fixed point O to a general point P of the curve, <strong>and</strong> p, theperpendicular distance from O to the tangent to the curve at P . By proceedingas indicated below, show that the radius of curvature, ρ, atP canbewritteninthe <strong>for</strong>m ρ = r dr/dp.Consider two neighbouring points, P <strong>and</strong> Q, on the curve. The normals to thecurve through those points meet at C, with (in the limit Q → P ) CP = CQ = ρ.Apply the cosine rule to triangles OPC <strong>and</strong> OQC to obtain two expressions <strong>for</strong>c 2 ,oneintermsofr <strong>and</strong> p <strong>and</strong> the other in terms of r +∆r <strong>and</strong> p +∆p. Byequating them <strong>and</strong> letting Q → P deduce the stated result.2.21 Use Leibnitz’ theorem to find(a) the second derivative of cos x sin 2x,(b) the third derivative of sin x ln x,(c) the fourth derivative of (2x 3 +3x 2 + x +2)exp2x.2.22 If y =exp(−x 2 ), show that dy/dx = −2xy <strong>and</strong> hence, by applying Leibnitz’theorem, prove that <strong>for</strong> n ≥ 1y (n+1) +2xy (n) +2ny (n−1) =0.2.23 Use the properties of functions at their turning points to do the following:(a) By considering its properties near x = 1, show that f(x) =5x 4 − 11x 3 +26x 2 − 44x + 24 takes negative values <strong>for</strong> some range of x.(b) Show that f(x) =tanx − x cannot be negative <strong>for</strong> 0 ≤ x


2.3 EXERCISES2.26 Use the mean value theorem to establish bounds in the following cases.(a) For − ln(1 − y), by considering ln x in the range 0 < 1 − y


PRELIMINARY CALCULUS(c) [(x − a)/(b − x)] 1/2 .2.38 Determine whether the following integrals exist <strong>and</strong>, where they do, evaluatethem: ∫ ∞∫ ∞x(a) exp(−λx) dx; (b)0−∞ (x 2 + a 2 ) dx;∫ 2 ∞∫11(c)1 x +1 dx; (d) 10 x dx;∫ 2 π/2∫ 1x(e) cot θdθ; (f)dx.00 (1 − x 2 )1/22.39 Use integration∫by parts to evaluate the following:y∫ y(a) x 2 sin xdx; (b) x ln xdx;(c)0∫ y0sin −1 xdx;(d)1∫ y1ln(a 2 + x 2 )/x 2 dx.2.40 Show, using the following methods, that the indefinite integral of x 3 /(x +1) 1/2 isJ = 235 (5x3 − 6x 2 +8x − 16)(x +1) 1/2 + c.(a) Repeated integration by parts.(b) Setting x +1=u 2 <strong>and</strong> determining dJ/du as (dJ/dx)(dx/du).2.41 The gamma function Γ(n) is defined <strong>for</strong> all n>−1 byΓ(n +1)=∫ ∞0x n e −x dx.Find a recurrence relation connecting Γ(n +1)<strong>and</strong>Γ(n).(a) Deduce (i) the value of Γ(n +1) when n is a non-negative integer, <strong>and</strong> (ii)the value of Γ ( ) (72 ,giventhatΓ 1) √2 = π.(b) Now, ( ) taking factorial m <strong>for</strong> any m to be defined by m! =Γ(m + 1), evaluate−32 !.2.42 Define J(m, n), <strong>for</strong> non-negative integers m <strong>and</strong> n, by the integralJ(m, n) =∫ π/20cos m θ sin n θdθ.(a) Evaluate J(0, 0), J(0, 1), J(1, 0), J(1, 1), J(m, 1), J(1,n).(b) Using integration by parts, prove that, <strong>for</strong> m <strong>and</strong> n both > 1,J(m, n) = m − 1n − 1J(m − 2,n) <strong>and</strong> J(m, n) = J(m, n − 2).m + n m + n(c) Evaluate (i) J(5, 3), (ii) J(6, 5) <strong>and</strong> (iii) J(4, 8).2.43 By integrating by parts twice, prove that I n as defined in the first equality below<strong>for</strong> positive integers n has the value given in the second equality:∫ π/2I n = sin nθ cos θdθ= n − sin(nπ/2) .0n 2 − 12.44 Evaluate the following definite integrals:∫ ∞[(x 3 +1)/(x 4 +4x +1) ] dx;(a)(c)0 xe−x dx; (b) ∫ 10∫ π/2[a +(a − 1) cos θ] −1 dθ with a> 1 ; (d) ∫ ∞0 2 −∞ (x2 +6x +18) −1 dx.80


2.4 HINTS AND ANSWERS2.45 If J r is the integral∫ ∞0x r exp(−x 2 ) dxshow that(a) J 2r+1 =(r!)/2,(b) J 2r =2 −r (2r − 1)(2r − 3) ···(5)(3)(1) J 0 .2.46 Find positive constants a, b such that ax ≤ sin x ≤ bx <strong>for</strong> 0 ≤ x ≤ π/2. Usethis inequality to find (to two significant figures) upper <strong>and</strong> lower bounds <strong>for</strong> theintegral∫ π/2I = (1+sinx) 1/2 dx.0Use the substitution t =tan(x/2) to evaluate I exactly.2.47 By noting that <strong>for</strong> 0 ≤ η ≤ 1, η 1/2 ≥ η 3/4 ≥ η, prove that23 ≤ 1 ∫ a(a 2 − x 2 ) 3/4 dx ≤ π a 5/2 04 .2.48 Show that the total length of the astroid x 2/3 + y 2/3 = a 2/3 , which can beparameterised as x = a cos 3 θ, y = a sin 3 θ,is6a.2.49 By noting that sinh x< 1 2 ex < cosh x, <strong>and</strong>that1+z 2 < (1 + z) 2 <strong>for</strong> z>0, showthat, <strong>for</strong> x>0, the length L of the curve y = 1 2 ex measured from the originsatisfies the inequalities sinh x


PRELIMINARY CALCULUSy2aπa2πaxFigure 2.14 The solution to exercise 2.17.2.21 (a) 2(2 − 9cos 2 x)sinx; (b)(2x −3 − 3x −1 )sinx − (3x −2 +lnx)cosx; (c)8(4x 3 +30x 2 +62x + 38) exp 2x.2.23 (a) f(1) = 0 whilst f ′ (1) ≠0,<strong>and</strong>sof(x) mustbenegativeinsomeregionwithx = 1 as an endpoint.(b) f ′ (x) =tan 2 x>0<strong>and</strong>f(0) = 0; g ′ (x) =(− cos x)(tan x − x)/x 2 , which isnever positive in the range.2.25 The false result arises because tan nx is not differentiable at x = π/(2n), whichlies in the range 0 0:(i) ∆ −1 ln[(2ax + b − ∆)/(2ax + b +∆)]+k;(ii) 2∆ ′−1 tan −1 [(2ax + b)/∆ ′ ]+k;(iii) −2(2ax + b) −1 + k.2.35 f ′ (x)=(1+sinx)/ cos 2 x = f(x)secx; J =ln(f(x)) + c =ln(secx +tanx)+c.2.37 Note that dx =2(b − a)cosθ sin θdθ.(a) π; (b)π(b − a) 2 /8; (c) π(b − a)/2.2.39 (a) (2 − y 2 )cosy + 2y sin y − 2; (b) [(y 2 ln y)/2] + [(1 − y 2 )/4];(c) y sin −1 y +(1− y 2 ) 1/2 − 1;(d) ln(a 2 +1)− (1/y)ln(a 2 + y 2 )+(2/a)[tan −1 (y/a) − tan −1 (1/a)].2.41 Γ(n +1)=nΓ(n); (a) (i) n!, (ii) 15 √ π/8; (b) −2 √ π.2.43 By integrating twice, recover a multiple of I n .2.45 J 2r+1 = rJ 2r−1 <strong>and</strong> 2J 2r =(2r − 1)J 2r−2 .2.47 Set η =1− (x/a) 2 throughout, <strong>and</strong> x = a sin θ in one of the bounds.2.49 L = ∫ x0(1+14 exp 2x) 1/2dx.82


3Complex numbers <strong>and</strong>hyperbolic functionsThis chapter is concerned with the representation <strong>and</strong> manipulation of complexnumbers. Complex numbers pervade this book, underscoring their wide applicationin the mathematics of the physical sciences. The application of complexnumbers to the description of physical systems is left until later chapters <strong>and</strong>only the basic tools are presented here.3.1 The need <strong>for</strong> complex numbersAlthough complex numbers occur in many branches of mathematics, they arisemost directly out of solving polynomial equations. We examine a specific quadraticequation as an example.Consider the quadratic equationEquation (3.1) has two solutions, z 1 <strong>and</strong> z 2 , such thatz 2 − 4z +5=0. (3.1)(z − z 1 )(z − z 2 )=0. (3.2)Using the familiar <strong>for</strong>mula <strong>for</strong> the roots of a quadratic equation, (1.4), thesolutions z 1 <strong>and</strong> z 2 , written in brief as z 1,2 ,arez 1,2 = 4 ± √ (−4) 2 − 4(1 × 5)√2−4=2±2 . (3.3)Both solutions contain the square root of a negative number. However, it is nottrue to say that there are no solutions to the quadratic equation. The fundamentaltheorem of algebra states that a quadratic equation will always have two solutions<strong>and</strong> these are in fact given by (3.3). The second term on the RHS of (3.3) iscalled an imaginary term since it contains the square root of a negative number;83


COMPLEX NUMBERS AND HYPERBOLIC FUNCTIONSf(z)543211234 zFigure 3.1 The function f(z) =z 2 − 4z +5.the first term is called a real term. The full solution is the sum of a real term<strong>and</strong> an imaginary term <strong>and</strong> is called a complex number. A plot of the functionf(z) =z 2 − 4z + 5 is shown in figure 3.1. It will be seen that the plot does notintersect the z-axis, corresponding to the fact that the equation f(z) =0hasnopurely real solutions.The choice of the symbol z <strong>for</strong> the quadratic variable was not arbitrary; theconventional representation of a complex number is z, wherez is the sum of areal part x <strong>and</strong> i times an imaginary part y, i.e.z = x + iy,where i is used to denote the square root of −1. The real part x <strong>and</strong> the imaginarypart y are usually denoted by Re z <strong>and</strong> Im z respectively. We note at this pointthat some physical scientists, engineers in particular, use j instead of i. However,<strong>for</strong> consistency, we will use i throughout this book.In our particular example, √ −4=2 √ −1=2i, <strong>and</strong> hence the two solutions of(3.1) arez 1,2 =2± 2i =2± i.2Thus, here x = 2 <strong>and</strong> y = ±1.For compactness a complex number is sometimes written in the <strong>for</strong>mz =(x, y),where the components of z may be thought of as coordinates in an xy-plot. Sucha plot is called an Arg<strong>and</strong> diagram <strong>and</strong> is a common representation of complexnumbers; an example is shown in figure 3.2.84


3.2 MANIPULATION OF COMPLEX NUMBERSIm zyz = x + iyxRe zFigure 3.2The Arg<strong>and</strong> diagram.Our particular example of a quadratic equation may be generalised readily topolynomials whose highest power (degree) is greater than 2, e.g. cubic equations(degree 3), quartic equations (degree 4) <strong>and</strong> so on. For a general polynomial f(z),of degree n, the fundamental theorem of algebra states that the equation f(z) =0will have exactly n solutions. We will examine cases of higher-degree equationsin subsection 3.4.3.The remainder of this chapter deals with: the algebra <strong>and</strong> manipulation ofcomplex numbers; their polar representation, which has advantages in manycircumstances; complex exponentials <strong>and</strong> logarithms; the use of complex numbersin finding the roots of polynomial equations; <strong>and</strong> hyperbolic functions.3.2 Manipulation of complex numbersThis section considers basic complex number manipulation. Some analogy maybe drawn with vector manipulation (see chapter 7) but this section st<strong>and</strong>s aloneas an introduction.3.2.1 Addition <strong>and</strong> subtractionThe addition of two complex numbers, z 1 <strong>and</strong> z 2 , in general gives anothercomplex number. The real components <strong>and</strong> the imaginary components are addedseparately <strong>and</strong> in a like manner to the familiar addition of real numbers:z 1 + z 2 =(x 1 + iy 1 )+(x 2 + iy 2 )=(x 1 + x 2 )+i(y 1 + y 2 ),85


COMPLEX NUMBERS AND HYPERBOLIC FUNCTIONSIm zz 2z 1 + z 2z 1Re zFigure 3.3The addition of two complex numbers.or in component notationz 1 + z 2 =(x 1 ,y 1 )+(x 2 ,y 2 )=(x 1 + x 2 ,y 1 + y 2 ).The Arg<strong>and</strong> representation of the addition of two complex numbers is shown infigure 3.3.By straight<strong>for</strong>ward application of the commutativity <strong>and</strong> associativity of thereal <strong>and</strong> imaginary parts separately, we can show that the addition of complexnumbers is itself commutative <strong>and</strong> associative, i.e.z 1 + z 2 = z 2 + z 1 ,z 1 +(z 2 + z 3 )=(z 1 + z 2 )+z 3 .Thus it is immaterial in what order complex numbers are added.◮Sum the complex numbers 1+2i, 3 − 4i, −2+i.Summing the real terms we obtain1+3− 2=2,<strong>and</strong> summing the imaginary terms we obtain2i − 4i + i = −i.Hence(1 + 2i)+(3− 4i)+(−2+i) =2− i. ◭The subtraction of complex numbers is very similar to their addition. As in thecase of real numbers, if two identical complex numbers are subtracted then theresult is zero.86


3.2 MANIPULATION OF COMPLEX NUMBERSIm zy|z|xarg zRe zFigure 3.4The modulus <strong>and</strong> argument of a complex number.3.2.2 Modulus <strong>and</strong> argumentThe modulus of the complex number z is denoted by |z| <strong>and</strong> is defined as|z| = √ x 2 + y 2 . (3.4)Hence the modulus of the complex number is the distance of the correspondingpoint from the origin in the Arg<strong>and</strong> diagram, as may be seen in figure 3.4.The argument of the complex number z is denoted by arg z <strong>and</strong> is defined asarg z =tan −1 ( yx). (3.5)It can be seen that arg z is the angle that the line joining the origin to z onthe Arg<strong>and</strong> diagram makes with the positive x-axis. The anticlockwise directionis taken to be positive by convention. The angle arg z is shown in figure 3.4.Account must be taken of the signs of x <strong>and</strong> y individually in determining inwhich quadrant arg z lies. Thus, <strong>for</strong> example, if x <strong>and</strong> y are both negative thenarg z lies in the range −π


COMPLEX NUMBERS AND HYPERBOLIC FUNCTIONS3.2.3 MultiplicationComplex numbers may be multiplied together <strong>and</strong> in general give a complexnumber as the result. The product of two complex numbers z 1 <strong>and</strong> z 2 is foundby multiplying them out in full <strong>and</strong> remembering that i 2 = −1, i.e.z 1 z 2 =(x 1 + iy 1 )(x 2 + iy 2 )= x 1 x 2 + ix 1 y 2 + iy 1 x 2 + i 2 y 1 y 2=(x 1 x 2 − y 1 y 2 )+i(x 1 y 2 + y 1 x 2 ). (3.6)◮Multiply the complex numbers z 1 =3+2i <strong>and</strong> z 2 = −1 − 4i.By direct multiplication we findz 1 z 2 =(3+2i)(−1 − 4i)= −3 − 2i − 12i − 8i 2=5− 14i. ◭ (3.7)The multiplication of complex numbers is both commutative <strong>and</strong> associative,i.e.z 1 z 2 = z 2 z 1 , (3.8)(z 1 z 2 )z 3 = z 1 (z 2 z 3 ). (3.9)The product of two complex numbers also has the simple propertiesThese relations are derived in subsection 3.3.1.|z 1 z 2 | = |z 1 ||z 2 |, (3.10)arg(z 1 z 2 )=argz 1 +arg z 2 . (3.11)◮Verify that (3.10) holds <strong>for</strong> the product of z 1 =3+2i <strong>and</strong> z 2 = −1 − 4i.From (3.7)We also find|z 1 z 2 | = |5 − 14i| = √ 5 2 +(−14) 2 = √ 221.|z 1 | = √ 3 2 +2 2 = √ 13,|z 2 | = √ (−1) 2 +(−4) 2 = √ 17,<strong>and</strong> hence|z 1 ||z 2 | = √ 13 √ 17 = √ 221 = |z 1 z 2 |. ◭We now examine the effect on a complex number z of multiplying it by ±1<strong>and</strong> ±i. These four multipliers have modulus unity <strong>and</strong> we can see immediatelyfrom (3.10) that multiplying z by another complex number of unit modulus givesa product with the same modulus as z. We can also see from (3.11) that if we88


3.2 MANIPULATION OF COMPLEX NUMBERSizIm zzRe z−z−izFigure 3.5Multiplication of a complex number by ±1 <strong>and</strong>±i.multiply z by a complex number then the argument of the product is the sumof the argument of z <strong>and</strong> the argument of the multiplier. Hence multiplyingz by unity (which has argument zero) leaves z unchanged in both modulus<strong>and</strong> argument, i.e. z is completely unaltered by the operation. Multiplying by−1 (which has argument π) leads to rotation, through an angle π, of the linejoining the origin to z in the Arg<strong>and</strong> diagram. Similarly, multiplication by i or −ileads to corresponding rotations of π/2 or−π/2 respectively. This geometricalinterpretation of multiplication is shown in figure 3.5.◮Using the geometrical interpretation of multiplication by i, find the product i(1 − i).The complex number 1 − i has argument −π/4 <strong>and</strong> modulus √ 2. Thus, using (3.10) <strong>and</strong>(3.11), its product with i has argument +π/4 <strong>and</strong> unchanged modulus √ 2. The complexnumber with modulus √ 2 <strong>and</strong> argument +π/4 is1+i <strong>and</strong> soi(1 − i) =1+i,as is easily verified by direct multiplication. ◭The division of two complex numbers is similar to their multiplication butrequires the notion of the complex conjugate (see the following subsection) <strong>and</strong>so discussion is postponed until subsection 3.2.5.3.2.4 Complex conjugateIf z has the convenient <strong>for</strong>m x + iy then the complex conjugate, denoted by z ∗ ,may be found simply by changing the sign of the imaginary part, i.e. if z = x + iythen z ∗ = x − iy. More generally, we may define the complex conjugate of z asthe (complex) number having the same magnitude as z that when multiplied byz leaves a real result, i.e. there is no imaginary component in the product.89


COMPLEX NUMBERS AND HYPERBOLIC FUNCTIONSIm zyz = x + iyxRe z−yz ∗ = x − iyFigure 3.6The complex conjugate as a mirror image in the real axis.Inthecasewherez can be written in the <strong>for</strong>m x + iy it is easily verified, bydirect multiplication of the components, that the product zz ∗ gives a real result:zz ∗ =(x + iy)(x − iy) =x 2 − ixy + ixy − i 2 y 2 = x 2 + y 2 = |z| 2 .Complex conjugation corresponds to a reflection of z in the real axis of theArg<strong>and</strong> diagram, as may be seen in figure 3.6.◮Find the complex conjugate of z = a +2i +3ib.The complex number is written in the st<strong>and</strong>ard <strong>for</strong>mz = a + i(2+3b);then, replacing i by −i, weobtainz ∗ = a − i(2+3b). ◭In some cases, however, it may not be simple to rearrange the expression <strong>for</strong>z into the st<strong>and</strong>ard <strong>for</strong>m x + iy. Nevertheless, given two complex numbers, z 1<strong>and</strong> z 2 , it is straight<strong>for</strong>ward to show that the complex conjugate of their sum(or difference) is equal to the sum (or difference) of their complex conjugates, i.e.(z 1 ± z 2 ) ∗ = z1 ∗ ± z∗ 2 . Similarly, it may be shown that the complex conjugate of theproduct (or quotient) of z 1 <strong>and</strong> z 2 is equal to the product (or quotient) of theircomplex conjugates, i.e. (z 1 z 2 ) ∗ = z1 ∗z∗ 2 <strong>and</strong> (z 1/z 2 ) ∗ = z1 ∗/z∗ 2 .Using these results, it can be deduced that, no matter how complicated theexpression, its complex conjugate may always be found by replacing every i by−i. To apply this rule, however, we must always ensure that all complex parts arefirst written out in full, so that no i’s are hidden.90


3.2 MANIPULATION OF COMPLEX NUMBERS◮Find the complex conjugate of the complex number z = w (3y+2ix) ,wherew = x +5i.Although we do not discuss complex powers until section 3.5, the simple rule given abovestill enables us to find the complex conjugate of z.In this case w itself contains real <strong>and</strong> imaginary components <strong>and</strong> so must be writtenout in full, i.e.z = w 3y+2ix =(x +5i) 3y+2ix .Now we can replace each i by −i to obtainz ∗ =(x − 5i) (3y−2ix) .It can be shown that the product zz ∗ is real, as required. ◭The following properties of the complex conjugate are easily proved <strong>and</strong> othersmay be derived from them. If z = x + iy then(z ∗ ) ∗ = z, (3.12)z + z ∗ =2Rez =2x, (3.13)z − z ∗ =2i Im z =2iy, (3.14)(z x 2z ∗ = − y 2 ) ( ) 2xyx 2 + y 2 + ix 2 + y 2 . (3.15)The derivation of this last relation relies on the results of the following subsection.3.2.5 DivisionThe division of two complex numbers z 1 <strong>and</strong> z 2 bears some similarity to theirmultiplication. Writing the quotient in component <strong>for</strong>m we obtainz 1= x 1 + iy 1. (3.16)z 2 x 2 + iy 2In order to separate the real <strong>and</strong> imaginary components of the quotient, wemultiply both numerator <strong>and</strong> denominator by the complex conjugate of thedenominator. By definition, this process will leave the denominator as a realquantity. Equation (3.16) givesz 1= (x 1 + iy 1 )(x 2 − iy 2 )z 2 (x 2 + iy 2 )(x 2 − iy 2 ) = (x 1x 2 + y 1 y 2 )+i(x 2 y 1 − x 1 y 2 )x 2 2 + y2 2= x 1x 2 + y 1 y 2x 2 2 + y2 2+ i x 2y 1 − x 1 y 2x 2 2 + .y2 2Hence we have separated the quotient into real <strong>and</strong> imaginary components, asrequired.In the special case where z 2 = z ∗ 1 ,sothatx 2 = x 1 <strong>and</strong> y 2 = −y 1 , the generalresult reduces to (3.15).91


COMPLEX NUMBERS AND HYPERBOLIC FUNCTIONS◮Express z in the <strong>for</strong>m x + iy, whenz =3 − 2i−1+4i .Multiplying numerator <strong>and</strong> denominator by the complex conjugate of the denominatorwe obtain(3 − 2i)(−1 − 4i) −11 − 10iz = =(−1+4i)(−1 − 4i) 17= − 1117 − 1017 i. ◭In analogy to (3.10) <strong>and</strong> (3.11), which describe the multiplication of twocomplex numbers, the following relations apply to division:∣ z 1 ∣∣∣∣ = |z 1|z 2 |z 2 | , (3.17)( )z1arg =argz 1 − arg z 2 . (3.18)z 2The proof of these relations is left until subsection 3.3.1.3.3 Polar representation of complex numbersAlthough considering a complex number as the sum of a real <strong>and</strong> an imaginarypart is often useful, sometimes the polar representation proves easier to manipulate.This makes use of the complex exponential function, which is defined bye z =expz ≡ 1+z + z22! + z3+ ··· . (3.19)3!Strictly speaking it is the function exp z that is defined by (3.19). The number eis the value of exp(1), i.e. it is just a number. However, it may be shown that e z<strong>and</strong> exp z are equivalent when z is real <strong>and</strong> rational <strong>and</strong> mathematicians thendefine their equivalence <strong>for</strong> irrational <strong>and</strong> complex z. For the purposes of thisbook we will not concern ourselves further with this mathematical nicety but,rather, assume that (3.19) is valid <strong>for</strong> all z. We also note that, using (3.19), bymultiplying together the appropriate series we may show that (see chapter 24)e z1 e z2 = e z1+z2 , (3.20)which is analogous to the familiar result <strong>for</strong> exponentials of real numbers.92


3.3 POLAR REPRESENTATION OF COMPLEX NUMBERSIm zyz = re iθrθxRe zFigure 3.7The polar representation of a complex number.From (3.19), it immediately follows that <strong>for</strong> z = iθ, θ real,e iθ =1+iθ − θ22! − iθ3 + ··· (3.21)3!(···)=1− θ22! + θ44! − ···+ i θ − θ33! + θ55! − (3.22)<strong>and</strong> hence thate iθ =cosθ + i sin θ, (3.23)where the last equality follows from the series expansions of the sine <strong>and</strong> cosinefunctions (see subsection 4.6.3). This last relationship is called Euler’s equation. Italso follows from (3.23) thate inθ =cosnθ + i sin nθ<strong>for</strong> all n. From Euler’s equation (3.23) <strong>and</strong> figure 3.7 we deduce thatre iθ = r(cos θ + i sin θ)= x + iy.Thus a complex number may be represented in the polar <strong>for</strong>mz = re iθ . (3.24)Referring again to figure 3.7, we can identify r with |z| <strong>and</strong> θ with arg z. Thesimplicity of the representation of the modulus <strong>and</strong> argument is one of the mainreasons <strong>for</strong> using the polar representation. The angle θ lies conventionally in therange −π


COMPLEX NUMBERS AND HYPERBOLIC FUNCTIONSIm zr 1 r 2 e i(θ 1+θ 2 )r 2 e iθ 2r 1 e iθ 1Re zFigure 3.8 The multiplication of two complex numbers. In this case r 1 <strong>and</strong>r 2 are both greater than unity.The algebra of the polar representation is different from that of the real <strong>and</strong>imaginary component representation, though, of course, the results are identical.Some operations prove much easier in the polar representation, others much morecomplicated. The best representation <strong>for</strong> a particular problem must be determinedby the manipulation required.3.3.1 Multiplication <strong>and</strong> division in polar <strong>for</strong>mMultiplication <strong>and</strong> division in polar <strong>for</strong>m are particularly simple. The product ofz 1 = r 1 e iθ1 <strong>and</strong> z 2 = r 2 e iθ2 is given byz 1 z 2 = r 1 e iθ1 r 2 e iθ2= r 1 r 2 e i(θ1+θ2) . (3.25)The relations |z 1 z 2 | = |z 1 ||z 2 | <strong>and</strong> arg(z 1 z 2 ) = arg z 1 +argz 2 follow immediately.An example of the multiplication of two complex numbers is shown in figure 3.8.Division is equally simple in polar <strong>for</strong>m; the quotient of z 1 <strong>and</strong> z 2 is given byz 1= r 1e iθ1z 2 r 2 e = r 1e i(θ1−θ2) . (3.26)iθ2 r 2The relations |z 1 /z 2 | = |z 1 |/|z 2 | <strong>and</strong> arg(z 1 /z 2 ) = arg z 1 − arg z 2 are again94


3.4 DE MOIVRE’S THEOREMIm zr 1 e iθ 1r 2 e iθ 2r 1e i(θ 1−θ 2 )r 2Re zFigure 3.9 The division of two complex numbers. As in the previous figure,r 1 <strong>and</strong> r 2 are both greater than unity.immediately apparent. The division of two complex numbers in polar <strong>for</strong>m isshown in figure 3.9.3.4 de Moivre’s theoremWe now derive an extremely important theorem. Since ( e iθ) n= e inθ , we have(cos θ + i sin θ) n =cosnθ + i sin nθ, (3.27)where the identity e inθ =cosnθ + i sin nθ follows from the series definition ofe inθ (see (3.21)). This result is called de Moivre’s theorem <strong>and</strong>isoftenusedinthemanipulation of complex numbers. The theorem is valid <strong>for</strong> all n whether real,imaginary or complex.There are numerous applications of de Moivre’s theorem but this sectionexamines just three: proofs of trigonometric identities; finding the nth roots ofunity; <strong>and</strong> solving polynomial equations with complex roots.3.4.1 Trigonometric identitiesThe use of de Moivre’s theorem in finding trigonometric identities is best illustratedby example. We consider the expression of a multiple-angle function interms of a polynomial in the single-angle function, <strong>and</strong> its converse.95


COMPLEX NUMBERS AND HYPERBOLIC FUNCTIONS◮Express sin 3θ <strong>and</strong> cos 3θ in terms of powers of cos θ <strong>and</strong> sin θ.Using de Moivre’s theorem,cos 3θ + i sin 3θ =(cosθ + i sin θ) 3=(cos 3 θ − 3cosθ sin 2 θ)+i(3 sin θ cos 2 θ − sin 3 θ). (3.28)We can equate the real <strong>and</strong> imaginary coefficients separately, i.e.<strong>and</strong>cos 3θ =cos 3 θ − 3cosθ sin 2 θ=4cos 3 θ − 3cosθ (3.29)sin 3θ =3sinθ cos 2 θ − sin 3 θ=3sinθ − 4sin 3 θ. ◭This method can clearly be applied to finding power expansions of cos nθ <strong>and</strong>sin nθ <strong>for</strong> any positive integer n.The converse process uses the following properties of z = e iθ ,z n + 1 =2cosnθ,zn (3.30)z n − 1 =2i sin nθ.zn (3.31)These equalities follow from simple applications of de Moivre’s theorem, i.e.z n + 1 z n =(cosθ + i sin θ)n +(cosθ + i sin θ) −n=cosnθ + i sin nθ +cos(−nθ)+i sin(−nθ)=cosnθ + i sin nθ +cosnθ − i sin nθ=2cosnθ<strong>and</strong>z n − 1 z n =(cosθ + i sin θ)n − (cos θ + i sin θ) −n=cosnθ + i sin nθ − cos nθ + i sin nθ=2i sin nθ.In the particular case where n =1,z + 1 z = eiθ + e −iθ =2cosθ, (3.32)z − 1 z = eiθ − e −iθ =2i sin θ. (3.33)96


3.4 DE MOIVRE’S THEOREM◮Find an expression <strong>for</strong> cos 3 θ in terms of cos 3θ <strong>and</strong> cos θ.Using (3.32),cos 3 θ = 1 (z + 1 ) 32 3 z(z 3 +3z + 3 z + 1 )z 3= 1 8= 1 8Now using (3.30) <strong>and</strong> (3.32), we find(z 3 + 1 )+ 3 (z + 1 ).z 3 8 zcos 3 θ = 1 cos 3θ + 3 cos θ. ◭4 4This result happens to be a simple rearrangement of (3.29), but cases involvinglarger values of n are better h<strong>and</strong>led using this direct method than by rearrangingpolynomial expansions of multiple-angle functions.3.4.2 Finding the nth roots of unityThe equation z 2 = 1 has the familiar solutions z = ±1. However, now thatwe have introduced the concept of complex numbers we can solve the generalequation z n = 1. Recalling the fundamental theorem of algebra, we know thatthe equation has n solutions. In order to proceed we rewrite the equation asz n = e 2ikπ ,where k is any integer. Now taking the nth root of each side of the equation wefindz = e 2ikπ/n .Hence, the solutions of z n = 1 arez 1,2,...,n =1, e 2iπ/n , ..., e 2i(n−1)π/n ,corresponding to the values 0, 1, 2,...,n− 1<strong>for</strong>k. Larger integer values of k donot give new solutions, since the roots already listed are simply cyclically repeated<strong>for</strong> k = n, n +1,n+2,etc.◮Find the solutions to the equation z 3 =1.By applying the above method we findz = e 2ikπ/3 .Hence the three solutions are z 1 = e 0i =1,z 2 = e 2iπ/3 , z 3 = e 4iπ/3 . We note that, as expected,the next solution, <strong>for</strong> which k =3,givesz 4 = e 6iπ/3 =1=z 1 , so that there are only threeseparate solutions. ◭97


COMPLEX NUMBERS AND HYPERBOLIC FUNCTIONSIm ze 2iπ/3 2π/32π/31Re ze −2iπ/3Figure 3.10 The solutions of z 3 =1.Not surprisingly, given that |z 3 | = |z| 3 from (3.10), all the roots of unity haveunit modulus, i.e. they all lie on a circle in the Arg<strong>and</strong> diagram of unit radius.The three roots are shown in figure 3.10.The cube roots of unity are often written 1, ω <strong>and</strong> ω 2 . The properties ω 3 =1<strong>and</strong> 1 + ω + ω 2 = 0 are easily proved.3.4.3 Solving polynomial equationsA third application of de Moivre’s theorem is to the solution of polynomialequations. Complex equations in the <strong>for</strong>m of a polynomial relationship must firstbe solved <strong>for</strong> z in a similar fashion to the method <strong>for</strong> finding the roots of realpolynomial equations. Then the complex roots of z may be found.◮Solve the equation z 6 − z 5 +4z 4 − 6z 3 +2z 2 − 8z +8=0.We first factorise to give(z 3 − 2)(z 2 +4)(z − 1) = 0.Hence z 3 =2orz 2 = −4 orz = 1. The solutions to the quadratic equation are z = ±2i;to find the complex cube roots, we first write the equation in the <strong>for</strong>mz 3 =2=2e 2ikπ ,where k is any integer. If we now take the cube root, we getz =2 1/3 e 2ikπ/3 .98


3.5 COMPLEX LOGARITHMS AND COMPLEX POWERSTo avoid the duplication of solutions, we use the fact that −π


COMPLEX NUMBERS AND HYPERBOLIC FUNCTIONSWe may use (3.34) to investigate further the properties of Ln z. We have alreadynoted that the argument of a complex number is multivalued, i.e. arg z = θ +2nπ,where n is any integer. Thus, in polar <strong>for</strong>m, the complex number z should strictlybe written asz = re i(θ+2nπ) .Taking the logarithm of both sides, <strong>and</strong> using (3.34), we findLn z =lnr + i(θ +2nπ), (3.35)where ln r is the natural logarithm of the real positive quantity r <strong>and</strong> so iswritten normally. Thus from (3.35) we see that Ln z is itself multivalued. To avoidthis multivalued behaviour it is conventional to define another function ln z, theprincipal value of Ln z, which is obtained from Ln z by restricting the argumentof z to lie in the range −π


3.6 APPLICATIONS TO DIFFERENTIATION AND INTEGRATION3.6 Applications to differentiation <strong>and</strong> integrationWe can use the exponential <strong>for</strong>m of a complex number together with de Moivre’stheorem (see section 3.4) to simplify the differentiation of trigonometric functions.◮Find the derivative with respect to x of e 3x cos 4x.We could differentiate this function straight<strong>for</strong>wardly using the product rule (see subsection2.1.2). However, an alternative method in this case is to use a complex exponential.Let us consider the complex numberz = e 3x (cos 4x + i sin 4x) =e 3x e 4ix = e (3+4i)x ,where we have used de Moivre’s theorem to rewrite the trigonometric functions as a complexexponential. This complex number has e 3x cos 4x as its real part. Now, differentiatingz with respect to x we obtaindzdx =(3+4i)e(3+4i)x =(3+4i)e 3x (cos 4x + i sin 4x), (3.36)where we have again used de Moivre’s theorem. Equating real parts we then findd (e 3x cos 4x ) = e 3x (3 cos 4x − 4sin4x).dxBy equating the imaginary parts of (3.36), we also obtain, as a bonus,d (e 3x sin 4x ) = e 3x (4 cos 4x +3sin4x). ◭dxIn a similar way the complex exponential can be used to evaluate integralscontaining trigonometric <strong>and</strong> exponential functions.◮Evaluate the integral I = ∫ e ax cos bx dx.Let us consider the integr<strong>and</strong> as the real part of the complex numbere ax (cos bx + i sin bx) =e ax e ibx = e (a+ib)x ,where we use de Moivre’s theorem to rewrite the trigonometric functions as a complexexponential. Integrating we find∫e (a+ib)x dx = e(a+ib)xa + ib + c(a − ib)e(a+ib)x=(a − ib)(a + ib) + c= eax (ae ibx − ibe ibx) + c, (3.37)a 2 + b 2where the constant of integration c is in general complex. Denoting this constant byc = c 1 + ic 2 <strong>and</strong> equating real parts in (3.37) we obtain∫I = e ax cos bx dx =eaxa 2 + b (a cos bx + b sin bx)+c 1,2which agrees with result (2.37) found using integration by parts. Equating imaginary partsin (3.37) we obtain, as a bonus,∫J = e ax sin bx dx =eaxa 2 + b (a sin bx − b cos bx)+c 2. ◭2101


COMPLEX NUMBERS AND HYPERBOLIC FUNCTIONS3.7 Hyperbolic functionsThe hyperbolic functions are the complex analogues of the trigonometric functions.The analogy may not be immediately apparent <strong>and</strong> their definitions may appearat first to be somewhat arbitrary. However, careful examination of their propertiesreveals the purpose of the definitions. For instance, their close relationship withthe trigonometric functions, both in their identities <strong>and</strong> in their calculus, meansthat many of the familiar properties of trigonometric functions can also be appliedto the hyperbolic functions. Further, hyperbolic functions occur regularly, <strong>and</strong> sogiving them special names is a notational convenience.3.7.1 DefinitionsThe two fundamental hyperbolic functions are cosh x <strong>and</strong> sinh x, which, as theirnames suggest, are the hyperbolic equivalents of cos x <strong>and</strong> sin x. They are definedby the following relations:cosh x = 1 2 (ex + e −x ), (3.38)sinh x = 1 2 (ex − e −x ). (3.39)Note that cosh x is an even function <strong>and</strong> sinh x is an odd function. By analogywith the trigonometric functions, the remaining hyperbolic functions aretanh x = sinh xcosh x = ex − e −xe x , (3.40)+ e−x sech x = 1cosh x = 2e x , (3.41)+ e−x cosech x = 1sinh x = 2e x ,− e−x (3.42)coth x = 1tanh x = ex + e −xe x .− e−x (3.43)All the hyperbolic functions above have been defined in terms of the real variablex. However, this was simply so that they may be plotted (see figures 3.11–3.13);the definitions are equally valid <strong>for</strong> any complex number z.3.7.2 Hyperbolic–trigonometric analogiesIn the previous subsections we have alluded to the analogy between trigonometric<strong>and</strong> hyperbolic functions. Here, we discuss the close relationship between the twogroups of functions.Recalling (3.32) <strong>and</strong> (3.33) we findcos ix = 1 2 (ex + e −x ),sin ix = 1 2 i(ex − e −x ).102


3.7 HYPERBOLIC FUNCTIONS43cosh x21sech x−2 −11 2xFigure 3.11Graphs of cosh x <strong>and</strong> sechx.4cosech x2sinh x−2 −11 2x−2cosech x−4Figure 3.12Graphs of sinh x <strong>and</strong> cosechx.Hence, by the definitions given in the previous subsection,cosh x =cosix, (3.44)i sinh x =sinix, (3.45)cos x =coshix, (3.46)i sin x = sinh ix. (3.47)These useful equations make the relationship between hyperbolic <strong>and</strong> trigono-103


COMPLEX NUMBERS AND HYPERBOLIC FUNCTIONS4coth x2tanh x−2 −11 2x−2coth x−4Figure 3.13 Graphs of tanh x <strong>and</strong> coth x.metric functions transparent. The similarity in their calculus is discussed furtherin subsection 3.7.6.3.7.3 Identities of hyperbolic functionsThe analogies between trigonometric functions <strong>and</strong> hyperbolic functions havingbeen established, we should not be surprised that all the trigonometric identitiesalso hold <strong>for</strong> hyperbolic functions, with the following modification. Whereversin 2 x occurs it must be replaced by − sinh 2 x, <strong>and</strong> vice versa. Note that thisreplacement is necessary even if the sin 2 x is hidden, e.g. tan 2 x =sin 2 x/ cos 2 x<strong>and</strong> so must be replaced by (− sinh 2 x/ cosh 2 x)=− tanh 2 x.◮Find the hyperbolic identity analogous to cos 2 x +sin 2 x =1.Using the rules stated above cos 2 x is replaced by cosh 2 x,<strong>and</strong>sin 2 x by − sinh 2 x,<strong>and</strong>sothe identity becomescosh 2 x − sinh 2 x =1.This can be verified by direct substitution, using the definitions of cosh x <strong>and</strong> sinh x; see(3.38) <strong>and</strong> (3.39). ◭Some other identities that can be proved in a similar way aresech 2 x =1− tanh 2 x, (3.48)cosech 2 x =coth 2 x − 1, (3.49)sinh 2x =2sinhx cosh x, (3.50)cosh 2x =cosh 2 x + sinh 2 x. (3.51)104


3.7 HYPERBOLIC FUNCTIONS3.7.4 Solving hyperbolic equationsWhen we are presented with a hyperbolic equation to solve, we may proceedby analogy with the solution of trigonometric equations. However, it is almostalways easier to express the equation directly in terms of exponentials.◮Solve the hyperbolic equation cosh x − 5sinhx − 5=0.Substituting the definitions of the hyperbolic functions we obtain12 (ex + e −x ) − 5 2 (ex − e −x ) − 5=0.Rearranging, <strong>and</strong> then multiplying through by −e x , gives in turn−2e x +3e −x − 5=0<strong>and</strong>2e 2x +5e x − 3=0.Now we can factorise <strong>and</strong> solve:(2e x − 1)(e x +3)=0.Thus e x =1/2 ore x = −3. Hence x = − ln 2 or x =ln(−3). The interpretation of thelogarithm of a negative number has been discussed in section 3.5. ◭3.7.5 Inverses of hyperbolic functionsJust like trigonometric functions, hyperbolic functions have inverses. If y =cosh x then x =cosh −1 y, which serves as a definition of the inverse. By usingthe fundamental definitions of hyperbolic functions, we can find closed-<strong>for</strong>mexpressions <strong>for</strong> their inverses. This is best illustrated by example.◮Find a closed-<strong>for</strong>m expression <strong>for</strong> the inverse hyperbolic function y =sinh −1 x.First we write x as a function of y, i.e.y =sinh −1 x ⇒ x =sinhy.Now, since cosh y = 1 2 (ey + e −y ) <strong>and</strong> sinh y = 1 2 (ey − e −y ),e y =coshy +sinhy√= 1+sinh 2 y +sinhye y = √ 1+x 2 + x,<strong>and</strong> hencey =ln( √ 1+x 2 + x). ◭In a similar fashion it can be shown thatcosh −1 x =ln( √ x 2 − 1+x).105


COMPLEX NUMBERS AND HYPERBOLIC FUNCTIONS42sech −1 xcosh −1 x123 4x−2−4sech −1 xcosh −1 xFigure 3.14 Graphs of cosh −1 x <strong>and</strong> sech −1 x.◮Find a closed-<strong>for</strong>m expression <strong>for</strong> the inverse hyperbolic function y =tanh −1 x.First we write x as a function of y, i.e.y =tanh −1 x ⇒ x =tanhy.Now, using the definition of tanh y <strong>and</strong> rearranging, we findx = ey − e −ye y + e −y ⇒ (x +1)e −y =(1− x)e y .Thus, it follows thate 2y = 1+x1 − x√1+x⇒ e y =1 − x ,√1+xy =ln1 − x ,tanh −1 x = 1 ( ) 1+x2 ln . ◭1 − xGraphs of the inverse hyperbolic functions are given in figures 3.14–3.16.3.7.6 Calculus of hyperbolic functionsJust as the identities of hyperbolic functions closely follow those of their trigonometriccounterparts, so their calculus is similar. The derivatives of the two basic106


3.7 HYPERBOLIC FUNCTIONS4cosech −1 x2sinh −1 x−2 −11 2xcosech −1 x−2−4Figure 3.15 Graphs of sinh −1 x <strong>and</strong> cosech −1 x.42tanh −1 xcoth −1 x−2 −11 2xcoth −1 x−2−4Figure 3.16 Graphs of tanh −1 x <strong>and</strong> coth −1 x.hyperbolic functions are given byd(cosh x) = sinh x, (3.52)dxd(sinh x) =coshx. (3.53)dxThey may be deduced by considering the definitions (3.38), (3.39) as follows.107


COMPLEX NUMBERS AND HYPERBOLIC FUNCTIONS◮Verify the relation (d/dx)coshx =sinhx.Using the definition of cosh x,cosh x = 1 2 (ex + e −x ),<strong>and</strong> differentiating directly, we findddx (cosh x) = 1 2 (ex − e −x )=sinhx. ◭Clearly the integrals of the fundamental hyperbolic functions are also definedby these relations. The derivatives of the remaining hyperbolic functions can bederived by product differentiation <strong>and</strong> are presented below only <strong>for</strong> completeness.ddx (tanh x) =sech2 x, (3.54)d(sech x) = −sech x tanh x,dx(3.55)d(cosech x) = −cosech x coth x,dx(3.56)ddx (coth x) = −cosech2 x. (3.57)The inverse hyperbolic functions also have derivatives, which are given by thefollowing:d(cosh −1 x )=dx ad(sinh −1 x )=dx ad(tanh −1 x )=dx addx(coth −1 x a1√x2 − a 2 , (3.58)1√x2 + a 2 , (3.59)aa 2 − x 2 , <strong>for</strong> x2 a 2 . (3.61)These may be derived from the logarithmic <strong>for</strong>m of the inverse (see subsection3.7.5).108


3.8 EXERCISES◮Evaluate (d/dx)sinh −1 x using the logarithmic <strong>for</strong>m of the inverse.From the results of section 3.7.5,d (sinh −1 x ) = ddxdx[ (ln1=x + √ x 2 +11=x + √ x 2 +1=1√x2 +1 . ◭x + √ )]x 2 +1(1+x√x2 +1)( √ )x2 +1+x√x2 +13.8 Exercises3.1 Two complex numbers z <strong>and</strong> w are given by z =3+4i <strong>and</strong> w =2− i. OnanArg<strong>and</strong> diagram, plot(a) z + w, (b)w − z, (c)wz, (d)z/w,(e) z ∗ w + w ∗ z,(f)w 2 ,(g)lnz, (h)(1+z + w) 1/2 .3.2 By considering the real <strong>and</strong> imaginary parts of the product e iθ e iφ prove thest<strong>and</strong>ard <strong>for</strong>mulae <strong>for</strong> cos(θ + φ) <strong>and</strong> sin(θ + φ).3.3 By writing π/12 = (π/3) − (π/4) <strong>and</strong> considering e iπ/12 , evaluate cot(π/12).3.4 Find the locus in the complex z-plane of points that satisfy the following equations.( ) 1+it(a) z − c = ρ , where c is complex, ρ is real <strong>and</strong> t is a real parameter1 − itthat varies in the range −∞


COMPLEX NUMBERS AND HYPERBOLIC FUNCTIONS(a) the equalities of corresponding angles, <strong>and</strong>(b) the constant ratio of corresponding sides,in the two triangles.By noting that any complex quantity can be expressed asz = |z| exp(i arg z),deduce thata(B − C)+b(C − A)+c(A − B) =0.3.9 For the real constant a find the loci of all points z = x + iy in the complex planethat satisfy{ ( )} z − ia(a) Re ln= c, c>0,z + ia{ ( )} z − ia(b) Im ln= k, 0 ≤ k ≤ π/2.z + iaIdentify the two families of curves <strong>and</strong> verify that in case (b) all curves passthrough the two points ±ia.3.10 The most general type of trans<strong>for</strong>mation between one Arg<strong>and</strong> diagram, in thez-plane, <strong>and</strong> another, in the Z-plane, that gives one <strong>and</strong> only one value of Z <strong>for</strong>each value of z (<strong>and</strong> conversely) is known as the general bilinear trans<strong>for</strong>mation<strong>and</strong> takes the <strong>for</strong>mz = aZ + bcZ + d .(a) Confirm that the trans<strong>for</strong>mation from the Z-plane to the z-plane is also ageneral bilinear trans<strong>for</strong>mation.(b) Recalling that the equation of a circle can be written in the <strong>for</strong>m∣ z − z 1 ∣∣∣∣ = λ, λ ≠1,z − z 2show that the general bilinear trans<strong>for</strong>mation trans<strong>for</strong>ms circles into circles(or straight lines). What is the condition that z 1 , z 2 <strong>and</strong> λ must satisfy if thetrans<strong>for</strong>med circle is to be a straight line?3.11 Sketch the parts of the Arg<strong>and</strong> diagram in which(a) Re z 2 < 0, |z 1/2 |≤2;(b) 0 ≤ arg z ∗ ≤ π/2;(c) | exp z 3 |→0as|z| →∞.What is the area of the region in which all three sets of conditions are satisfied?3.12 Denote the nth roots of unity by 1, ω n , ωn, 2 ... , ωn n−1 .(a) Prove that∑n−1∏n−1(i) ωn r =0, (ii) ωn r =(−1) n+1 .r=0(b) Express x 2 + y 2 + z 2 − yz − zx − xy as the product of two factors, each linearin x, y <strong>and</strong> z, with coefficients dependent on the third roots of unity (<strong>and</strong>those of the x terms arbitrarily taken as real).r=0110


3.8 EXERCISES3.13 Prove that x 2m+1 − a 2m+1 ,wherem is an integer ≥ 1, can be written asm∏[( ) ]2πrx 2m+1 − a 2m+1 =(x − a) x 2 − 2ax cos+ a 2 .2m +1r=13.14 The complex position vectors of two parallel interacting equal fluid vorticesmoving with their axes of rotation always perpendicular to the z-plane are z 1<strong>and</strong> z 2 . The equations governing their motions aredz ∗ 1dt= − iz 1 − z 2,dz ∗ 2dt= − iz 2 − z 1.Deduce that (a) z 1 + z 2 ,(b)|z 1 − z 2 | <strong>and</strong> (c) |z 1 | 2 + |z 2 | 2 are all constant in time,<strong>and</strong> hence describe the motion geometrically.3.15 Solve the equationz 7 − 4z 6 +6z 5 − 6z 4 +6z 3 − 12z 2 +8z +4=0,(a) by examining the effect of setting z 3 equal to 2, <strong>and</strong> then(b) by factorising <strong>and</strong> using the binomial expansion of (z + a) 4 .Plot the seven roots of the equation on an Arg<strong>and</strong> plot, exemplifying that complexroots of a polynomial equation always occur in conjugate pairs if the polynomialhas real coefficients.3.16 The polynomial f(z) is defined byf(z) =z 5 − 6z 4 +15z 3 − 34z 2 +36z − 48.(a) Show that the equation f(z) = 0 has roots of the <strong>for</strong>m z = λi, whereλ isreal, <strong>and</strong> hence factorize f(z).(b) Show further that the cubic factor of f(z) can be written in the <strong>for</strong>m(z + a) 3 + b, wherea <strong>and</strong> b are real, <strong>and</strong> hence solve the equation f(z) =0completely.3.17 The binomial expansion of (1 + x) n , discussed in chapter 1, can be written <strong>for</strong> apositive integer n asn∑(1 + x) n =n C r x r ,where n C r = n!/[r!(n − r)!].(a) Use de Moivre’s theorem to show that the sumS 1 (n) = n C 0 − n C 2 + n C 4 − ···+(−1) m n C 2m , n− 1 ≤ 2m ≤ n,has the value 2 n/2 cos(nπ/4).(b) Derive a similar result <strong>for</strong> the sumS 2 (n) = n C 1 − n C 3 + n C 5 − ···+(−1) m n C 2m+1 , n− 1 ≤ 2m +1≤ n,<strong>and</strong> verify it <strong>for</strong> the cases n =6,7<strong>and</strong>8.3.18 By considering (1 + exp iθ) n , prove thatn∑n C r cos rθ =2 n cos n (θ/2) cos(nθ/2),r=0r=0n∑n C r sin rθ =2 n cos n (θ/2) sin(nθ/2),r=0where n C r = n!/[r!(n − r)!].111


COMPLEX NUMBERS AND HYPERBOLIC FUNCTIONS3.19 Use de Moivre’s theorem with n = 4 to prove thatcos 4θ =8cos 4 θ − 8cos 2 θ +1,<strong>and</strong> deduce that(cos π 8 = 2+ √ ) 1/22.43.20 Express sin 4 θ entirely in terms of the trigonometric functions of multiple angles<strong>and</strong> deduce that its average value over a complete cycle is 3 . 83.21 Use de Moivre’s theorem to prove thattan 5θ = t5 − 10t 3 +5t5t 4 − 10t 2 +1 ,where t =tanθ. Deduce the values of tan(nπ/10) <strong>for</strong> n =1,2,3,4.3.22 Prove the following results involving hyperbolic functions.(a) That( ) x + ycosh x − cosh y =2sinh sinh2(b) That, if y =sinh −1 x,(x 2 +1) d2 ydx + x dy2 dx =0.3.23 Determine the conditions under which the equationa cosh x + b sinh x = c, c > 0,( x − yhas zero, one, or two real solutions <strong>for</strong> x. What is the solution if a 2 = c 2 + b 2 ?3.24 Use the definitions <strong>and</strong> properties of hyperbolic functions to do the following:(a) Solve cosh x =sinhx +2sechx.(b) Show that the real solution x of tanh x =cosechx canbewritteninthe<strong>for</strong>m x =ln(u + √ u). Find an explicit value <strong>for</strong> u.(c) Evaluate tanh x when x is the real solution of cosh 2x =2coshx.3.25 Express sinh 4 x in terms of hyperbolic cosines of multiples of x, <strong>and</strong> hence findthe real solutions of2cosh4x − 8cosh2x +5=0.2).3.26 In the theory of special relativity, the relationship between the position <strong>and</strong> timecoordinates of an event, as measured in two frames of reference that have parallelx-axes, can be expressed in terms of hyperbolic functions. If the coordinates arex <strong>and</strong> t in one frame <strong>and</strong> x ′ <strong>and</strong> t ′ in the other, then the relationship take the<strong>for</strong>mx ′ = x cosh φ − ct sinh φ,ct ′ = −x sinh φ + ct cosh φ.Express x <strong>and</strong> ct in terms of x ′ , ct ′ <strong>and</strong> φ <strong>and</strong> show thatx 2 − (ct) 2 =(x ′ ) 2 − (ct ′ ) 2 .112


3.9 HINTS AND ANSWERS3.27 A closed barrel has as its curved surface the surface obtained by rotating aboutthe x-axis the part of the curvey = a[2 − cosh(x/a)]lying in the range −b ≤ x ≤ b, whereb0. With these conditions,there are two roots if a 2 >b 2 , but only one if b 2 >a 2 .For a 2 = c 2 + b 2 , x = 1 ln[(a − b)/(a + b)].23.25 Reduce the equation to 16 sinh 4 x = 1, yielding x = ±0.481.113


COMPLEX NUMBERS AND HYPERBOLIC FUNCTIONS3.27 Show that ds =(coshx/a) dx;curved surface area = πa 2 [8 sinh(b/a) − sinh(2b/a)] − 2πab.flat ends area = 2πa 2 [4 − 4cosh(b/a)+cosh 2 (b/a)].114


4Series <strong>and</strong> limits4.1 SeriesMany examples exist in the physical sciences of situations where we are presentedwith a sum of terms to evaluate. For example, we may wish to add the contributionsfrom successive slits in a diffraction grating to find the total light intensity at aparticular point behind the grating.A series may have either a finite or infinite number of terms. In either case, thesum of the first N terms of a series (often called a partial sum) is writtenS N = u 1 + u 2 + u 3 + ···+ u N ,where the terms of the series u n , n =1, 2, 3,...,N are numbers, that may ingeneral be complex. If the terms are complex then S N will in general be complexalso, <strong>and</strong> we can write S N = X N + iY N ,whereX N <strong>and</strong> Y N are the partial sums ofthe real <strong>and</strong> imaginary parts of each term separately <strong>and</strong> are there<strong>for</strong>e real. If aseries has only N terms then the partial sum S N is of course the sum of the series.Sometimes we may encounter series where each term depends on some variable,x, say. In this case the partial sum of the series will depend on the value assumedby x. For example, consider the infinite seriesS(x) =1+x + x22! + x33! + ··· .This is an example of a power series; these are discussed in more detail insection 4.5. It is in fact the Maclaurin expansion of exp x (see subsection 4.6.3).There<strong>for</strong>e S(x) =expx <strong>and</strong>, of course, varies according to the value of thevariable x. A series might just as easily depend on a complex variable z.A general, r<strong>and</strong>om sequence of numbers can be described as a series <strong>and</strong> a sumof the terms found. However, <strong>for</strong> cases of practical interest, there will usually be115


SERIES AND LIMITSsome sort of relationship between successive terms. For example, if the nth termof a series is given byu n = 1 2 n ,<strong>for</strong> n =1, 2, 3,...,N then the sum of the first N terms will beS N =N∑n=1u n = 1 2 + 1 4 + 1 8 + ···+ 12 N . (4.1)It is clear that the sum of a finite number of terms is always finite, providedthat each term is itself finite. It is often of practical interest, however, to considerthe sum of a series with an infinite number of finite terms. The sum of aninfinite number of terms is best defined by first considering the partial sumof the first N terms, S N . If the value of the partial sum S N tends to a finitelimit, S, asN tends to infinity, then the series is said to converge <strong>and</strong> its sumis given by the limit S. In other words, the sum of an infinite series is givenbyS = limN→∞S N ,provided the limit exists. For complex infinite series, if S N approaches a limitS = X + iY as N →∞, this means that X N → X <strong>and</strong> Y N → Y separately, i.e.the real <strong>and</strong> imaginary parts of the series are each convergent series with sumsX <strong>and</strong> Y respectively.However, not all infinite series have finite sums. As N →∞, the value of thepartial sum S N may diverge: it may approach +∞ or −∞, or oscillate finitelyor infinitely. Moreover, <strong>for</strong> a series where each term depends on some variable,its convergence can depend on the value assumed by the variable. Whether aninfinite series converges, diverges or oscillates has important implications whendescribing physical systems. <strong>Methods</strong> <strong>for</strong> determining whether a series convergesare discussed in section 4.3.4.2 Summation of seriesIt is often necessary to find the sum of a finite series or a convergent infiniteseries. We now describe arithmetic, geometric <strong>and</strong> arithmetico-geometric series,which are particularly common <strong>and</strong> <strong>for</strong> which the sums are easily found. Othermethods that can sometimes be used to sum more complicated series are discussedbelow.116


4.2 SUMMATION OF SERIES4.2.1 Arithmetic seriesAn arithmetic series has the characteristic that the difference between successiveterms is constant. The sum of a general arithmetic series is writtenN−1∑S N = a +(a + d)+(a +2d)+···+ [a +(N − 1)d] = (a + nd).Rewriting the series in the opposite order <strong>and</strong> adding this term by term to theoriginal expression <strong>for</strong> S N , we findS N = N 2 [a + a +(N − 1)d] = N (first term + last term). (4.2)2If an infinite number of such terms are added the series will increase (or decrease)indefinitely; that is to say, it diverges.n=0◮Sum the integers between 1 <strong>and</strong> 1000 inclusive.This is an arithmetic series with a =1,d =1<strong>and</strong>N = 1000. There<strong>for</strong>e, using (4.2) we findS N = 1000 (1 + 1000) = 500500,2which can be checked directly only with considerable ef<strong>for</strong>t. ◭4.2.2 Geometric seriesEquation (4.1) is a particular example of a geometric series, which has thecharacteristic that the ratio of successive terms is a constant (one-half in thiscase). The sum of a geometric series is in general writtenN−1∑S N = a + ar + ar 2 + ···+ ar N−1 = ar n ,where a is a constant <strong>and</strong> r is the ratio of successive terms, the common ratio. Thesum may be evaluated by considering S N <strong>and</strong> rS N :n=0S N = a + ar + ar 2 + ar 3 + ···+ ar N−1 ,rS N = ar + ar 2 + ar 3 + ar 4 + ···+ ar N .If we now subtract the second equation from the first we obtain<strong>and</strong> hence(1 − r)S N = a − ar N ,S N = a(1 − rN ). (4.3)1 − r117


SERIES AND LIMITSFor a series with an infinite number of terms <strong>and</strong> |r| < 1, we have lim N→∞ r N =0,<strong>and</strong> the sum tends to the limitS =a1 − r . (4.4)In (4.1), r = 1 2 , a = 1 2,<strong>and</strong>soS =1.For|r| ≥1, however, the series either divergesor oscillates.◮Consider a ball that drops from a height of 27 m <strong>and</strong> on each bounce retains only a thirdof its kinetic energy; thus after one bounce it will return to a height of 9m,aftertwobounces to 3m, <strong>and</strong> so on. Find the total distance travelled between the first bounce <strong>and</strong>the Mth bounce.The total distance travelled between the first bounce <strong>and</strong> the Mth bounce is given by thesum of M − 1 terms:M−2∑ 9S M−1 =2(9+3+1+···) =23 m<strong>for</strong> M>1, where the factor 2 is included to allow <strong>for</strong> both the upward <strong>and</strong> the downwardjourney. Inside the parentheses we clearly have a geometric series with first term 9 <strong>and</strong>common ratio 1/3 <strong>and</strong> hence the distance is given by (4.3), i.e.[9 1 − ( ) ]1 M−1[3S M−1 =2×=27 1 − ( ) ]1 M−1,1 − 1 33where the number of terms N in (4.3) has been replaced by M − 1. ◭m=04.2.3 Arithmetico-geometric seriesAn arithmetico-geometric series, as its name suggests, is a combined arithmetic<strong>and</strong> geometric series. It has the general <strong>for</strong>mN−1∑S N = a +(a + d)r +(a +2d)r 2 + ···+ [a +(N − 1)d] r N−1 = (a + nd)r n ,<strong>and</strong> can be summed, in a similar way to a pure geometric series, by multiplyingby r <strong>and</strong> subtracting the result from the original series to obtain(1 − r)S N = a + rd + r 2 d + ···+ r N−1 d − [a +(N − 1)d] r N .Using the expression <strong>for</strong> the sum of a geometric series (4.3) <strong>and</strong> rearranging, wefindS N =a − [a +(N − 1)d] rN1 − r+ rd(1 − rN−1 )(1 − r) 2 .For an infinite series with |r| < 1, lim N→∞ r N = 0 as in the previous subsection,<strong>and</strong> the sum tends to the limitS =a1 − r + rd(1 − r) 2 . (4.5)As <strong>for</strong> a geometric series, if |r| ≥1 then the series either diverges or oscillates.118n=0


4.2 SUMMATION OF SERIES◮Sum the seriesS =2+ 5 2 + 8 2 2 + 112 3 + ··· .This is an infinite arithmetico-geometric series with a =2,d =3<strong>and</strong>r =1/2. There<strong>for</strong>e,from (4.5), we obtain S = 10. ◭4.2.4 The difference methodThe difference method is sometimes useful in summing series that are morecomplicated than the examples discussed above. Let us consider the general seriesN∑u n = u 1 + u 2 + ···+ u N .n=1If the terms of the series, u n , can be expressed in the <strong>for</strong>mu n = f(n) − f(n − 1)<strong>for</strong> some function f(n) then its (partial) sum is given byN∑S N = u n = f(N) − f(0).n=1This can be shown as follows. The sum is given byS N = u 1 + u 2 + ···+ u N<strong>and</strong> since u n = f(n) − f(n − 1), it may be rewrittenS N =[f(1) − f(0)] + [f(2) − f(1)] + ···+[f(N) − f(N − 1)].By cancelling terms we see thatS N = f(N) − f(0).◮Evaluate the sumN∑n=11n(n +1) .Using partial fractions we find( 1u n = −n +1 − 1 ).nHence u n = f(n) − f(n − 1) with f(n) =−1/(n +1),<strong>and</strong>sothesumisgivenbyS N = f(N) − f(0) = − 1N +1 +1= NN +1 . ◭119


SERIES AND LIMITSThe difference method may be easily extended to evaluate sums in which eachterm can be expressed in the <strong>for</strong>mu n = f(n) − f(n − m), (4.6)where m is an integer. By writing out the sum to N terms with each term expressedin this <strong>for</strong>m, <strong>and</strong> cancelling terms in pairs as be<strong>for</strong>e, we findm∑m∑S N = f(N − k +1)− f(1 − k).k=1k=1◮Evaluate the sumN∑n=11n(n +2) .Using partial fractions we find[u n = −12(n +2) − 1 2nHence u n = f(n) − f(n − 2) with f(n) =−1/[2(n +2)],<strong>and</strong>sothesumisgivenbyS N = f(N)+f(N − 1) − f(0) − f(−1) = 3 4 − 1 ( 12 N +2 + 1 ). ◭N +1In fact the difference method is quite flexible <strong>and</strong> may be used to evaluatesums even when each term cannot be expressed as in (4.6). The method still relies,however, on being able to write u n in terms of a single function such that mostterms in the sum cancel, leaving only a few terms at the beginning <strong>and</strong> the end.This is best illustrated by an example.].◮Evaluate the sumN∑n=11n(n +1)(n +2) .Using partial fractions we find1u n =2(n +2) − 1n +1 + 1 2n .Hence u n = f(n) − 2f(n − 1) + f(n − 2) with f(n) =1/[2(n + 2)]. If we write out the sum,expressing each term u n in this <strong>for</strong>m, we find that most terms cancel <strong>and</strong> the sum is givenbyS N = f(N) − f(N − 1) − f(0) + f(−1) = 1 4 + 1 2( 1N +2 − 1 ). ◭N +1120


4.2 SUMMATION OF SERIES4.2.5 Series involving natural numbersSeries consisting of the natural numbers 1, 2, 3, ..., or the square or cube of thesenumbers, occur frequently <strong>and</strong> deserve a special mention. Let us first considerthe sum of the first N natural numbers,N∑S N =1+2+3+···+ N = n.This is clearly an arithmetic series with first term a = 1 <strong>and</strong> common differenced = 1. There<strong>for</strong>e, from (4.2), S N = 1 2N(N +1).Next, we consider the sum of the squares of the first N natural numbers:N∑S N =1 2 +2 2 +3 2 + ...+ N 2 = n 2 ,which may be evaluated using the difference method. The nth term in the seriesis u n = n 2 , which we need to express in the <strong>for</strong>m f(n) − f(n − 1) <strong>for</strong> some functionf(n). Consider the functionf(n) =n(n + 1)(2n +1) ⇒ f(n − 1) = (n − 1)n(2n − 1).For this function f(n) − f(n − 1) = 6n 2 , <strong>and</strong> so we can writeu n = 1 6[f(n) − f(n − 1)].There<strong>for</strong>e, by the difference method,S N = 1 6 [f(N) − f(0)] = 1 6N(N + 1)(2N +1).Finally, we calculate the sum of the cubes of the first N natural numbers,N∑S N =1 3 +2 3 +3 3 + ···+ N 3 = n 3 ,again using the difference method. Consider the functionf(n) =[n(n +1)] 2 ⇒ f(n − 1) = [(n − 1)n] 2 ,<strong>for</strong> which f(n) − f(n − 1) = 4n 3 . There<strong>for</strong>e we can write the general nth term ofthe series asu n = 1 4[f(n) − f(n − 1)],<strong>and</strong> using the difference method we findS N = 1 4 [f(N) − f(0)] = 1 4 N2 (N +1) 2 .Note that this is the square of the sum of the natural numbers, i.e.(N∑N 2∑n 3 = n).n=1121n=1n=1n=1n=1


SERIES AND LIMITS◮Sum the seriesN∑(n +1)(n +3).n=1The nth term in this series isu n =(n +1)(n +3)=n 2 +4n +3,<strong>and</strong> there<strong>for</strong>e we can writeN∑N∑(n +1)(n +3)= (n 2 +4n +3)n=1n=1N∑N∑ N∑= n 2 +4 n + 3n=1n=1n=1= 1 N(N + 1)(2N +1)+4× 1 N(N +1)+3N6 2= 1 6 N(2N2 +15N + 31). ◭4.2.6 Trans<strong>for</strong>mation of seriesA complicated series may sometimes be summed by trans<strong>for</strong>ming it into afamiliar series <strong>for</strong> which we already know the sum, perhaps a geometric seriesor the Maclaurin expansion of a simple function (see subsection 4.6.3). Varioustechniques are useful, <strong>and</strong> deciding which one to use in any given case is a matterof experience. We now discuss a few of the more common methods.The differentiation or integration of a series is often useful in trans<strong>for</strong>ming anapparently intractable series into a more familiar one. If we wish to differentiateor integrate a series that already depends on some variable then we may do soin a straight<strong>for</strong>ward manner.◮Sum the seriesS(x) =x43(0!) + x54(1!) + x65(2!) + ··· .Dividing both sides by x we obtainS(x)x = x33(0!) + x44(1!) + x55(2!) + ··· ,which is easily differentiated to give[ ]d S(x)= x2dx x 0! + x31! + x42! + x53! + ··· .Recalling the Maclaurin expansion of exp x given in subsection 4.6.3, we recognise thatthe RHS is equal to x 2 exp x. Having done so, we can now integrate both sides to obtain∫S(x)/x = x 2 exp xdx.122


4.2 SUMMATION OF SERIESIntegrating the RHS by parts we findS(x)/x = x 2 exp x − 2x exp x +2expx + c,where the value of the constant of integration c canbefixedbytherequirementthatS(x)/x =0atx = 0. Thus we find that c = −2 <strong>and</strong> that the sum is given byS(x) =x 3 exp x − 2x 2 exp x +2x exp x − 2x. ◭Often, however, we require the sum of a series that does not depend on avariable. In this case, in order that we may differentiate or integrate the series,we define a function of some variable x such that the value of this function isequal to the sum of the series <strong>for</strong> some particular value of x (usually at x =1).◮Sum the seriesS =1+ 2 2 + 3 2 2 + 4 2 3 + ··· .Let us begin by defining the functionf(x)=1+2x +3x 2 +4x 3 + ··· ,so that the sum S = f(1/2). Integrating this function we obtain∫f(x) dx = x + x 2 + x 3 + ··· ,which we recognise as an infinite geometric series with first term a = x <strong>and</strong> common ratior = x. There<strong>for</strong>e, from (4.4), we find that the sum of this series is x/(1 − x). In other words∫f(x) dx =x1 − x ,so that f(x) isgivenbyf(x) = d ( x)1=dx 1 − x (1 − x) . 2The sum of the original series is there<strong>for</strong>e S = f(1/2) = 4. ◭Aside from differentiation <strong>and</strong> integration, an appropriate substitution cansometimes trans<strong>for</strong>m a series into a more familiar <strong>for</strong>m. In particular, series withterms that contain trigonometric functions can often be summed by the use ofcomplex exponentials.◮Sum the seriesS(θ)=1+cosθ +cos 2θ2!+cos 3θ3!+ ··· .Replacing the cosine terms with a complex exponential, we obtain{}exp 2iθ exp 3iθS(θ) =Re 1+expiθ + + + ···2! 3!}(exp iθ)2 (exp iθ)3=Re{1+expiθ + + + ··· .2! 3!123


SERIES AND LIMITSAgain using the Maclaurin expansion of exp x given in subsection 4.6.3, we notice thatS(θ) = Re [exp(exp iθ)] = Re [exp(cos θ + i sin θ)]=Re{[exp(cos θ)][exp(i sin θ)]} = [exp(cos θ)]Re [exp(i sin θ)]=[exp(cosθ)][cos(sin θ)]. ◭4.3 Convergence of infinite seriesAlthough the sums of some commonly occurring infinite series may be found,the sum of a general infinite series is usually difficult to calculate. Nevertheless,it is often useful to know whether the partial sum of such a series converges toa limit, even if the limit cannot be found explicitly. As mentioned at the end ofsection 4.1, if we allow N to tend to infinity, the partial sumN∑S N =u nn=1of a series may tend to a definite limit (i.e. the sum S of the series), or increaseor decrease without limit, or oscillate finitely or infinitely.To investigate the convergence of any given series, it is useful to have availablea number of tests <strong>and</strong> theorems of general applicability. We discuss them below;some we will merely state, since once they have been stated they become almostself-evident, but are no less useful <strong>for</strong> that.4.3.1 Absolute <strong>and</strong> conditional convergenceLet us first consider some general points concerning the convergence, or otherwise,of an infinite series. In general an infinite series ∑ u n can have complex terms,<strong>and</strong> in the special case of a real series the terms can be positive or negative. Fromany such series, however, we can always construct another series ∑ |u n | in whicheach term is simply the modulus of the corresponding term in the original series.Then each term in the new series will be a positive real number.If the series ∑ |u n | converges then ∑ u n also converges, <strong>and</strong> ∑ u n is said to beabsolutely convergent, i.e. the series <strong>for</strong>med by the absolute values is convergent.For an absolutely convergent series, the terms may be reordered without affectingthe convergence of the series. However, if ∑ |u n | diverges whilst ∑ u n convergesthen ∑ u n is said to be conditionally convergent. For a conditionally convergentseries, rearranging the order of the terms can affect the behaviour of the sum<strong>and</strong>, hence, whether the series converges or diverges. In fact, a theorem dueto Riemann shows that, by a suitable rearrangement, a conditionally convergentseries may be made to converge to any arbitrary limit, or to diverge, or to oscillatefinitely or infinitely! Of course, if the original series ∑ u n consists only of positivereal terms <strong>and</strong> converges then automatically it is absolutely convergent.124


4.3 CONVERGENCE OF INFINITE SERIES4.3.2 Convergence of a series containing only real positive termsAs discussed above, in order to test <strong>for</strong> the absolute convergence of a series∑un , we first construct the corresponding series ∑ |u n | that consists only of realpositive terms. There<strong>for</strong>e in this subsection we will restrict our attention to seriesof this type.We discuss below some tests that may be used to investigate the convergence ofsuch a series. Be<strong>for</strong>e doing so, however, we note the following crucial consideration.In all the tests <strong>for</strong>, or discussions of, the convergence of a series, it is not whathappens in the first ten, or the first thous<strong>and</strong>, or the first million terms (or anyother finite number of terms) that matters, but what happens ultimately.Preliminary testA necessary but not sufficient condition <strong>for</strong> a series of real positive terms ∑ u nto be convergent is that the term u n tends to zero as n tends to infinity, i.e. werequirelim u n =0.n→∞If this condition is not satisfied then the series must diverge. Even if it is satisfied,however, the series may still diverge, <strong>and</strong> further testing is required.Comparison testThe comparison test is the most basic test <strong>for</strong> convergence. Let us consider twoseries ∑ u n <strong>and</strong> ∑ v n <strong>and</strong> suppose that we know the latter to be convergent (bysome earlier analysis, <strong>for</strong> example). Then, if each term u n in the first series is lessthan or equal to the corresponding term v n in the second series, <strong>for</strong> all n greaterthan some fixed number N that will vary from series to series, then the originalseries ∑ u n is also convergent. In other words, if ∑ v n is convergent <strong>and</strong>u n ≤ v n<strong>for</strong> n>N,then ∑ u n converges.However, if ∑ v n diverges <strong>and</strong> u n ≥ v n <strong>for</strong> all n greater than some fixed numberthen ∑ u n diverges.◮Determine whether the following series converges:∞∑ 1n!+1 = 1 2 + 1 3 + 1 7 + 1 + ··· . (4.7)25n=1Let us compare this series with the series∞∑ 1n! = 1 0! + 1 1! + 1 2! + 1 3! + ···=2+ 1 2! + 1 + ··· , (4.8)3!n=0125


SERIES AND LIMITSwhich is merely the series obtained by setting x = 1 in the Maclaurin expansion of exp x(see subsection 4.6.3), i.e.exp(1) = e =1+ 1 1! + 1 2! + 1 3! + ··· .Clearly this second series is convergent, since it consists of only positive terms <strong>and</strong> has afinite sum. Thus, since each term u n in the series (4.7) is less than the corresponding term1/n! in (4.8), we conclude from the comparison test that (4.7) is also convergent. ◭D’Alembert’s ratio testThe ratio test determines whether a series converges by comparing the relativemagnitude of successive terms. If we consider a series ∑ u n <strong>and</strong> set( )un+1ρ = lim , (4.9)n→∞ u nthen if ρ1 the series is divergent; if ρ =1then the behaviour of the series is undetermined by this test.To prove this we observe that if the limit (4.9) is less than unity, i.e. ρ


4.3 CONVERGENCE OF INFINITE SERIESRatio comparison testAs its name suggests, the ratio comparison test is a combination of the ratio <strong>and</strong>comparison tests. Let us consider the two series ∑ u n <strong>and</strong> ∑ v n <strong>and</strong> assume thatwe know the latter to be convergent. It may be shown that ifu n+1u n≤ v n+1v n<strong>for</strong> all n greater than some fixed value N then ∑ u n is also convergent.Similarly, ifu n+1≥ v n+1u n v n<strong>for</strong> all sufficiently large n, <strong>and</strong> ∑ v n diverges then ∑ u n also diverges.◮Determine whether the following series converges:∞∑ 1(n!) =1+ 1 2 2 + 1 2 6 + ··· .2n=1In this case the ratio of successive terms, as n tends to infinity, is given by[ ] 2 ( ) 2n!1R = lim= lim ,n→∞ (n +1)! n→∞ n +1which is less than the ratio seen in (4.11). Hence, by the ratio comparison test, the seriesconverges. (It is clear that this series could also be found to be convergent using the ratiotest.) ◭Quotient testThe quotient test may also be considered as a combination of the ratio <strong>and</strong>comparison tests. Let us again consider the two series ∑ u n <strong>and</strong> ∑ v n , <strong>and</strong> defineρ as the limit( )unρ = lim . (4.12)n→∞ v nThen, it can be shown that:(i) if ρ ≠ 0 but is finite then ∑ u n <strong>and</strong> ∑ v n either both converge or bothdiverge;(ii) if ρ = 0 <strong>and</strong> ∑ v n converges then ∑ u n converges;(iii) if ρ = ∞ <strong>and</strong> ∑ v n diverges then ∑ u n diverges.127


SERIES AND LIMITS◮Given that the series ∑ ∞n=11/n diverges, determine whether the following series converges:∞∑ 4n 2 − n − 3n 3 +2n . (4.13)n=1If we set u n =(4n 2 − n − 3)/(n 3 +2n) <strong>and</strong>v n =1/n then the limit (4.12) becomes[ ] [ ](4n 2 − n − 3)/(n 3 +2n) 4n 3 − n 2 − 3nρ = lim= lim=4.n→∞ 1/nn→∞ n 3 +2nSince ρ is finite but non-zero <strong>and</strong> ∑ v n diverges, from (i) above ∑ u n must also diverge. ◭Integral testThe integral test is an extremely powerful means of investigating the convergenceof a series ∑ u n . Suppose that there exists a function f(x) which monotonicallydecreases <strong>for</strong> x greater than some fixed value x 0 <strong>and</strong> <strong>for</strong> which f(n) =u n ,i.e.thevalue of the function at integer values of x is equal to the corresponding termin the series under investigation. Then it can be shown that, if the limit of theintegral∫ Nlim f(x) dxN→∞exists, the series ∑ u n is convergent. Otherwise the series diverges. Note that theintegral defined here has no lower limit; the test is sometimes stated with a lowerlimit, equal to unity, <strong>for</strong> the integral, but this can lead to unnecessary difficulties.◮Determine whether the following series converges:∞∑ 1(n − 3/2) =4+4+ 4 2 9 + 425 + ··· .n=1Let us consider the function f(x) =(x − 3/2) −2 . Clearly f(n) =u n <strong>and</strong> f(x) monotonicallydecreases <strong>for</strong> x>3/2. Applying the integral test, we consider∫ N( )1−1limdx = lim=0.N→∞ (x − 3/2)2 N→∞ N − 3/2Since the limit exists the series converges. Note, however, that if we had included a lowerlimit, equal to unity, in the integral then we would have run into problems, since theintegr<strong>and</strong> diverges at x =3/2. ◭The integral test is also useful <strong>for</strong> examining the convergence of the Riemannzeta series. This is a special series that occurs regularly <strong>and</strong> is of the <strong>for</strong>m∞∑ 1n p .n=1It converges <strong>for</strong> p>1 <strong>and</strong> diverges if p ≤ 1. These convergence criteria may bederived as follows.128


4.3 CONVERGENCE OF INFINITE SERIESUsing the integral test, we consider∫ N( )1N1−plim dx = lim ,N→∞ xp N→∞ 1 − p<strong>and</strong> it is obvious that the limit tends to zero <strong>for</strong> p>1<strong>and</strong>to∞ <strong>for</strong> p ≤ 1.Cauchy’s root testCauchy’s root test may be useful in testing <strong>for</strong> convergence, especially if the nthterms of the series contains an nth power. If we define the limitρ = lim (u n ) 1/n ,n→∞then it may be proved that the series ∑ u n converges if ρ1 then theseries diverges. Its behaviour is undetermined if ρ =1.◮Determine whether the following series converges:∞∑( ) n 1=1+ 1 n 4 + 1 27 + ··· .n=1Using Cauchy’s root test, we find<strong>and</strong> hence the series converges. ◭( ) 1ρ = lim =0,n→∞ nGrouping termsWe now consider the Riemann zeta series, mentioned above, with an alternativeproof of its convergence that uses the method of grouping terms. In general thereare better ways of determining convergence, but the grouping method may beused if it is not immediately obvious how to approach a problem by a bettermethod.First consider the case where p>1, <strong>and</strong> group the terms in the series asfollows:S N = 1 ( 11 p + 2 p + 1 ) ( 13 p +4 p + ···+ 1 )7 p + ··· .Now we can see that each bracket of this series is less than each term of thegeometric seriesS N = 1 1 p + 2 2 p + 4 4 p + ··· .This geometric series has common ratio r = ( 12) p−1;sincep>1, it follows thatr1.129


SERIES AND LIMITSThe divergence of the Riemann zeta series <strong>for</strong> p ≤ 1 can be seen by firstconsidering the case p = 1. The series isS N =1+ 1 2 + 1 3 + 1 4 + ··· ,which does not converge, as may be seen by bracketing the terms of the series ingroups in the following way:N∑( ( 1 1S N = u n =1+ +2)3 4)+ 1 ( 1+5 + 1 6 + 1 7 8)+ 1 + ··· .n=1The sum of the terms in each bracket is ≥ 1 2<strong>and</strong>, since as many such groupingscan be made as we wish, it is clear that S N increases indefinitely as N is increased.Now returning to the case of the Riemann zeta series <strong>for</strong> p 1/n <strong>for</strong> n>1, p


4.4 OPERATIONS WITH SERIESis always less than u N <strong>for</strong> all m <strong>and</strong> u n → 0asn →∞, the alternating seriesconverges. It is clear that an analogous proof can be constructed in the casewhere N is even.◮Determine whether the following series converges:∞∑(−1) n+1 1 n =1− 1 2 + 1 3 − ··· .n=1This alternating series clearly satisfies conditions (i) <strong>and</strong> (ii) above <strong>and</strong> hence converges.However, as shown above by the method of grouping terms, the corresponding series withall positive terms is divergent. ◭4.4 Operations with seriesSimple operations with series are fairly intuitive, <strong>and</strong> we discuss them here only<strong>for</strong> completeness. The following points apply to both finite <strong>and</strong> infinite seriesunless otherwise stated.(i) If ∑ u n = S then ∑ ku n = kS where k is any constant.(ii) If ∑ u n = S <strong>and</strong> ∑ v n = T then ∑ (u n + v n )=S + T .(iii) If ∑ u n = S then a + ∑ u n = a + S. A simple extension of this trivial resultshows that the removal or insertion of a finite number of terms anywherein a series does not affect its convergence.(iv) If the infinite series ∑ u n <strong>and</strong> ∑ v n are both absolutely convergent thenthe series ∑ w n ,wherew n = u 1 v n + u 2 v n−1 + ···+ u n v 1 ,is also absolutely convergent. The series ∑ w n is called the Cauchy productof the two original series. Furthermore, if ∑ u n converges to the sum S<strong>and</strong> ∑ v n converges to the sum T then ∑ w n converges to the sum ST.(v) It is not true in general that term-by-term differentiation or integration ofa series will result in a new series with the same convergence properties.A power series has the <strong>for</strong>m4.5 Power seriesP (x) =a 0 + a 1 x + a 2 x 2 + a 3 x 3 + ··· ,where a 0 ,a 1 ,a 2 ,a 3 etc. are constants. Such series regularly occur in physics <strong>and</strong>engineering <strong>and</strong> are useful because, <strong>for</strong> |x| < 1, the later terms in the series maybecome very small <strong>and</strong> be discarded. For example the seriesP (x) =1+x + x 2 + x 3 + ··· ,131


SERIES AND LIMITSalthough in principle infinitely long, in practice may be simplified if x happens tohave a value small compared with unity. To see this note that P (x) <strong>for</strong>x =0.1has the following values: 1, if just one term is taken into account; 1.1, <strong>for</strong> twoterms; 1.11, <strong>for</strong> three terms; 1.111, <strong>for</strong> four terms, etc. If the quantity that itrepresents can only be measured with an accuracy of two decimal places, then allbut the first three terms may be ignored, i.e. when x =0.1 orlessP (x) =1+x + x 2 +O(x 3 ) ≈ 1+x + x 2 .This sort of approximation is often used to simplify equations into manageable<strong>for</strong>ms. It may seem imprecise at first but is perfectly acceptable insofar as itmatches the experimental accuracy that can be achieved.The symbols O <strong>and</strong> ≈ used above need some further explanation. They areused to compare the behaviour of two functions when a variable upon whichboth functions depend tends to a particular limit, usually zero or infinity (<strong>and</strong>obvious from the context). For two functions f(x) <strong>and</strong>g(x), with g positive, the<strong>for</strong>mal definitions of the above symbols are as follows:(i) If there exists a constant k such that |f| ≤kg as the limit is approachedthen f =O(g).(ii) If as the limit of x is approached f/g tends to a limit l, wherel ≠0,thenf ≈ lg. The statement f ≈ g means that the ratio of the two sides tendsto unity.4.5.1 Convergence of power seriesThe convergence or otherwise of power series is a crucial consideration in practicalterms. For example, if we are to use a power series as an approximation, it isclearly important that it tends to the precise answer as more <strong>and</strong> more terms ofthe approximation are taken. Consider the general power seriesP (x) =a 0 + a 1 x + a 2 x 2 + ··· .Using d’Alembert’s ratio test (see subsection 4.3.2), we see that P (x) convergesabsolutely if∣ ∣ ∣∣∣ a n+1ρ = lim xn→∞ a n∣ = |x| lima n+1 ∣∣∣n→∞ ∣ < 1.a nThus the convergence of P (x) depends upon the value of x, i.e. there is, in general,a range of values of x <strong>for</strong> which P (x) converges, an interval of convergence. Notethat at the limits of this range ρ = 1, <strong>and</strong> so the series may converge or diverge.The convergence of the series at the end-points may be determined by substitutingthese values of x into the power series P (x) <strong>and</strong> testing the resulting series usingany applicable method (discussed in section 4.3).132


4.5 POWER SERIES◮Determine the range of values of x <strong>for</strong> which the following power series converges:P (x)=1+2x +4x 2 +8x 3 + ··· .By using the interval-of-convergence method discussed above,∣ ∣ ∣∣∣ 2 n+1 ∣∣∣ρ = limn→∞ 2 x = |2x|,n<strong>and</strong> hence the power series will converge <strong>for</strong> |x| < 1/2. Examining the end-points of theinterval separately, we findP (1/2)=1+1+1+··· ,P (−1/2) = 1 − 1+1− ··· .Obviously P (1/2) diverges, while P (−1/2) oscillates. There<strong>for</strong>e P (x) is not convergent ateither end-point of the region but is convergent <strong>for</strong> −1


SERIES AND LIMITSr = − exp iθ. There<strong>for</strong>e, on the the circle of convergence we have1P (z) =1+expiθ .Unless θ = π this is a finite complex number, <strong>and</strong> so P (z) converges at all points on thecircle |z| =2exceptatθ = π (i.e. z = −2), where it diverges. Note that P (z) isjustthebinomial expansion of (1 + z/2) −1 , <strong>for</strong> which it is obvious that z = −2 is a singular point.In general, <strong>for</strong> power series expansions of complex functions about a given point in thecomplex plane, the circle of convergence extends as far as the nearest singular point. Thisis discussed further in chapter 24. ◭Note that the centre of the circle of convergence does not necessarily lie at theorigin. For example, applying the ratio test to the complex power seriesP (z) =1+ z − 12+(z − 1)24+(z − 1)38+ ··· ,we find that <strong>for</strong> it to converge we require |(z − 1)/2| < 1. Thus the series converges<strong>for</strong> z lying within a circle of radius 2 centred on the point (1, 0) in the Arg<strong>and</strong>diagram.4.5.2 Operations with power seriesThe following rules are useful when manipulating power series; they apply topower series in a real or complex variable.(i) If two power series P (x) <strong>and</strong>Q(x) have regions of convergence that overlapto some extent then the series produced by taking the sum, the difference or theproduct of P (x) <strong>and</strong>Q(x) converges in the common region.(ii) If two power series P (x) <strong>and</strong>Q(x) converge <strong>for</strong> all values of x then oneseries may be substituted into the other to give a third series, which also converges<strong>for</strong> all values of x. For example, consider the power series expansions of sin x <strong>and</strong>e x given below in subsection 4.6.3,sin x = x − x33! + x55! − x77! + ···e x =1+x + x22! + x33! + x44! + ··· ,both of which converge <strong>for</strong> all values of x. Substituting the series <strong>for</strong> sin x intothat <strong>for</strong> e x we obtaine sin x =1+x + x22! − 3x44! − 8x55! + ··· ,which also converges <strong>for</strong> all values of x.If, however, either of the power series P (x) <strong>and</strong>Q(x) has only a limited regionof convergence, or if they both do so, then further care must be taken whensubstituting one series into the other. For example, suppose Q(x) converges <strong>for</strong>all x, but P (x) onlyconverges<strong>for</strong>x within a finite range. We may substitute134


4.5 POWER SERIESQ(x) intoP (x) toobtainP (Q(x)), but we must be careful since the value of Q(x)may lie outside the region of convergence <strong>for</strong> P (x), with the consequence that theresulting series P (Q(x)) does not converge.(iii) If a power series P (x) converges <strong>for</strong> a particular range of x then the seriesobtained by differentiating every term <strong>and</strong> the series obtained by integrating everyterm also converge in this range.This is easily seen <strong>for</strong> the power seriesP (x) =a 0 + a 1 x + a 2 x 2 + ··· ,which converges if |x| < lim n→∞ |a n /a n+1 |≡k. The series obtained by differentiatingP (x) with respect to x is given bydPdx = a 1 +2a 2 x +3a 3 x 2 + ···<strong>and</strong> converges if∣ ∣ ∣∣∣ na n ∣∣∣|x| < lim= k.n→∞ (n +1)a n+1Similarly the series obtained by integrating P (x) term by term,converges if∫P (x) dx = a 0 x + a 1x 22+ a 2x 33+ ··· ,∣ ∣ ∣∣∣ (n +2)a n ∣∣∣|x| < lim= k.n→∞ (n +1)a n+1So, series resulting from differentiation or integration have the same interval ofconvergence as the original series. However, even if the original series convergesat either end-point of the interval, it is not necessarily the case that the new serieswill do so. The new series must be tested separately at the end-points in orderto determine whether it converges there. Note that although power series may beintegrated or differentiated without altering their interval of convergence, this isnot true <strong>for</strong> series in general.It is also worth noting that differentiating or integrating a power series termby term within its interval of convergence is equivalent to differentiating orintegrating the function it represents. For example, consider the power seriesexpansion of sin x,sin x = x − x33! + x55! − x7+ ··· , (4.14)7!which converges <strong>for</strong> all values of x. If we differentiate term by term, the seriesbecomes1 − x22! + x44! − x66! + ··· ,which is the series expansion of cos x, asweexpect.135


SERIES AND LIMITS4.6 Taylor seriesTaylor’s theorem provides a way of expressing a function as a power series in x,known as a Taylor series, but it can be applied only to those functions that arecontinuous <strong>and</strong> differentiable within the x-range of interest.4.6.1 Taylor’s theoremSuppose that we have a function f(x) that we wish to express as a power seriesin x − a about the point x = a. We shall assume that, in a given x-range, f(x)is a continuous, single-valued function of x having continuous derivatives withrespect to x, denoted by f ′ (x), f ′′ (x) <strong>and</strong> so on, up to <strong>and</strong> including f (n−1) (x). Weshall also assume that f (n) (x) exists in this range.From the equation following (2.31) we may write∫ a+haf ′ (x) dx = f(a + h) − f(a),where a, a + h are neighbouring values of x. Rearranging this equation, we mayexpress the value of the function at x = a + h in terms of its value at a byf(a + h) =f(a)+∫ a+haf ′ (x) dx. (4.15)A first approximation <strong>for</strong> f(a + h) may be obtained by substituting f ′ (a) <strong>for</strong>f ′ (x) in (4.15), to obtainf(a + h) ≈ f(a)+hf ′ (a).This approximation is shown graphically in figure 4.1. We may write this firstapproximation in terms of x <strong>and</strong> a as<strong>and</strong>, in a similar way,f(x) ≈ f(a)+(x − a)f ′ (a),f ′ (x) ≈ f ′ (a)+(x − a)f ′′ (a),f ′′ (x) ≈ f ′′ (a)+(x − a)f ′′′ (a),<strong>and</strong> so on. Substituting <strong>for</strong> f ′ (x) in (4.15), we obtain the second approximation:f(a + h) ≈ f(a)+∫ a+ha≈ f(a)+hf ′ (a)+ h22 f′′ (a).[f ′ (a)+(x − a)f ′′ (a)] dxWe may repeat this procedure as often as we like (so long as the derivativesof f(x) exist) to obtain higher-order approximations to f(a + h); we find the136


4.6 TAYLOR SERIESf(x)QRf(a)Pθhhf ′ (a)aa + hxFigure 4.1 The first-order Taylor series approximation to a function f(x).The slope of the function at P ,i.e.tanθ, equalsf ′ (a). Thus the value of thefunction at Q, f(a + h), is approximated by the ordinate of R, f(a)+hf ′ (a).(n − 1)th-order approximation § to bef(a + h) ≈ f(a)+hf ′ (a)+ h22! f′′ (a)+···+ hn−1(n − 1)! f(n−1) (a). (4.16)As might have been anticipated, the error associated with approximating f(a+h)by this (n − 1)th-order power series is of the order of the next term in the series.This error or remainder can be shown to be given byR n (h) = hnn! f(n) (ξ),<strong>for</strong> some ξ that lies in the range [a, a + h]. Taylor’s theorem then states that wemay write the equalityf(a + h) =f(a)+hf ′ (a)+ h22! f′′ (a)+···+ h(n−1)(n − 1)! f(n−1) (a)+R n (h).(4.17)The theorem may also be written in a <strong>for</strong>m suitable <strong>for</strong> finding f(x) giventhe value of the function <strong>and</strong> its relevant derivatives at x = a, by substituting§ The order of the approximation is simply the highest power of h in the series. Note, though, thatthe (n − 1)th-order approximation contains n terms.137


SERIES AND LIMITSx = a + h in the above expression. It then readsf(x) =f(a)+(x − a)f ′ (a)+where the remainder now takes the <strong>for</strong>m(x − a)2f ′′ (x − a)n−1(a)+···+2!(n − 1)! f(n−1) (a)+R n (x),(4.18)(x − a)nR n (x) = f (n) (ξ),n!<strong>and</strong> ξ lies in the range [a, x]. Each of the <strong>for</strong>mulae (4.17), (4.18) gives us theTaylor expansion of the function about the point x = a. A special case occurswhen a = 0. Such Taylor expansions, about x = 0, are called Maclaurin series.Taylor’s theorem is also valid without significant modification <strong>for</strong> functionsof a complex variable (see chapter 24). The extension of Taylor’s theorem tofunctions of more than one variable is given in chapter 5.For a function to be expressible as an infinite power series we require it to beinfinitely differentiable <strong>and</strong> the remainder term R n to tend to zero as n tends toinfinity, i.e. lim n→∞ R n = 0. In this case the infinite power series will represent thefunction within the interval of convergence of the series.◮Exp<strong>and</strong> f(x) =sinx as a Maclaurin series, i.e. about x =0.We must first verify that sin x may indeed be represented by an infinite power series. It iseasily shown that the nth derivative of f(x) isgivenby(f (n) (x) =sin x + nπ ).2There<strong>for</strong>e the remainder after exp<strong>and</strong>ing f(x) asan(n − 1)th-order polynomial aboutx = 0 is given byR n (x) =(ξ xnn! sin + nπ ),2where ξ lies in the range [0,x]. Since the modulus of the sine term is always less than orequal to unity, we can write |R n (x)| < |x n |/n!. For any particular value of x, sayx = c,R n (c) → 0asn →∞. Hence lim n→∞ R n (x) = 0, <strong>and</strong> so sin x can be represented by aninfinite Maclaurin series.Evaluating the function <strong>and</strong> its derivatives at x =0weobtainf(0) = sin 0 = 0,f ′ (0) = sin(π/2) = 1,f ′′ (0) = sin π =0,f ′′′ (0) = sin(3π/2) = −1,<strong>and</strong> so on. There<strong>for</strong>e, the Maclaurin series expansion of sin x is given bysin x = x − x33! + x55! − ··· .Note that, as expected, since sin x is an odd function, its power series expansion containsonly odd powers of x. ◭138


4.6 TAYLOR SERIESWe may follow a similar procedure to obtain a Taylor series about an arbitrarypoint x = a.◮Exp<strong>and</strong> f(x) =cosx as a Taylor series about x = π/3.As in the above example, it is easily shown that the nth derivative of f(x) isgivenby(f (n) (x) =cos x + nπ ).2There<strong>for</strong>e the remainder after exp<strong>and</strong>ing f(x) asan(n − 1)th-order polynomial aboutx = π/3 isgivenbyR n (x) =(x − π/3)nn!(cos ξ + nπ ),2where ξ lies in the range [π/3,x]. The modulus of the cosine term is always less than orequal to unity, <strong>and</strong> so |R n (x)| < |(x−π/3) n |/n!. As in the previous example, lim n→∞ R n (x) =0 <strong>for</strong> any particular value of x, <strong>and</strong>socosx can be represented by an infinite Taylor seriesabout x = π/3.Evaluating the function <strong>and</strong> its derivatives at x = π/3 weobtainf(π/3) = cos(π/3) = 1/2,f ′ (π/3) = cos(5π/6) = − √ 3/2,f ′′ (π/3) = cos(4π/3) = −1/2,<strong>and</strong> so on. Thus the Taylor series expansion of cos x about x = π/3 isgivenbycos x = 1 2 − √32(x − π/3)−12( ) 2x − π/3+ ··· . ◭2!4.6.2 Approximation errors in Taylor seriesIn the previous subsection we saw how to represent a function f(x) by an infinitepower series, which is exactly equal to f(x) <strong>for</strong> all x within the interval ofconvergence of the series. However, in physical problems we usually do not wantto have to sum an infinite number of terms, but prefer to use only a finite numberof terms in the Taylor series to approximate the function in some given rangeof x. In this case it is desirable to know what is the maximum possible errorassociated with the approximation.As given in (4.18), a function f(x) can be represented by a finite (n − 1)th-orderpower series together with a remainder term such thatf(x) =f(a)+(x − a)f ′ (a)+(x − a)2f ′′ (x − a)n−1(a)+···+2!(n − 1)! f(n−1) (a)+R n (x),where(x − a)nR n (x) = f (n) (ξ)n!<strong>and</strong> ξ lies in the range [a, x]. R n (x) is the remainder term, <strong>and</strong> represents the errorin approximating f(x) bytheabove(n − 1)th-order power series. Since the exact139


SERIES AND LIMITSvalue of ξ that satisfies the expression <strong>for</strong> R n (x) is not known, an upper limit onthe error may be found by differentiating R n (x) withrespecttoξ <strong>and</strong> equatingthe derivative to zero in the usual way <strong>for</strong> finding maxima.◮Exp<strong>and</strong> f(x) =cosx as a Taylor series about x =0<strong>and</strong> find the error associated withusing the approximation to evaluate cos(0.5) if only the first two non-vanishing terms aretaken. (Note that the Taylor expansions of trigonometric functions are only valid <strong>for</strong> anglesmeasured in radians.)Evaluating the function <strong>and</strong> its derivatives at x = 0, we findf(0) = cos 0 = 1,f ′ (0) = − sin 0 = 0,f ′′ (0) = − cos 0 = −1,f ′′′ (0) = sin 0 = 0.So, <strong>for</strong> small |x|, we find from (4.18)cos x ≈ 1 − x22 .Note that since cos x is an even function, its power series expansion contains only evenpowers of x. There<strong>for</strong>e, in order to estimate the error in this approximation, we mustconsider the term in x 4 , which is the next in the series. The required derivative is f (4) (x)<strong>and</strong> this is (by chance) equal to cos x. Thus, adding in the remainder term R 4 (x), we findcos x =1− x22 + x4cos ξ,4!where ξ lies in the range [0,x]. Thus, the maximum possible error is x 4 /4!, since cos ξcannot exceed unity. If x =0.5, taking just the first two terms yields cos(0.5) ≈ 0.875 witha predicted error of less than 0.00260. In fact cos(0.5) = 0.87758 to 5 decimal places. Thus,to this accuracy, the true error is 0.00258, an error of about 0.3%. ◭4.6.3 St<strong>and</strong>ard Maclaurin seriesIt is often useful to have a readily available table of Maclaurin series <strong>for</strong> st<strong>and</strong>ardelementary functions, <strong>and</strong> there<strong>for</strong>e these are listed below.sin x = x − x33! + x55! − x7+ ···7!<strong>for</strong> −∞


4.7 EVALUATION OF LIMITSThese can all be derived by straight<strong>for</strong>ward application of Taylor’s theorem tothe expansion of a function about x =0.4.7 Evaluation of limitsThe idea of the limit of a function f(x) asx approaches a value a is fairly intuitive,though a strict definition exists <strong>and</strong> is stated below. In many cases the limit ofthe function as x approaches a will be simply the value f(a), but sometimes thisis not so. Firstly, the function may be undefined at x = a, as, <strong>for</strong> example, whenf(x) = sin xx ,which takes the value 0/0 atx = 0. However, the limit as x approaches zerodoes exist <strong>and</strong> can be evaluated as unity using l’Hôpital’s rule below. Anotherpossibility is that even if f(x) is defined at x = a its value may not be equal to thelimiting value lim x→a f(x). This can occur <strong>for</strong> a discontinuous function at a pointof discontinuity. The strict definition of a limit is that if lim x→a f(x) =l then<strong>for</strong> any number ɛ however small, it must be possible to find a number η such that|f(x)−l|


SERIES AND LIMITS◮Evaluate the limitslimx→1 (x2 +2x 3 ),lim(x cos x),x→0sin xlimx→π/2 x .Using (a) above,Using (b),Using (c),limx→1 (x2 +2x 3 ) = lim x 2 + lim 2x 3 =3.x→1 x→1lim(x cos x) = lim x lim cos x =0× 1=0.x→0 x→0 x→0sin xlimx→π/2 x= lim x→π/2 sin xlim x→π/2 x = 1π/2 = 2 π . ◭(iv) Limits of functions of x that contain exponents that themselves depend onx can often be found by taking logarithms.◮Evaluate the limit) x 2lim(1 − a2.x→∞ x 2Let us definelim ln y = limx→∞ x→∞) x 2y =(1 − a2x 2<strong>and</strong> consider the logarithm of the required limit, i.e.[x 2 ln(1 − a2x 2 )].Using the Maclaurin series <strong>for</strong> ln(1 + x) given in subsection 4.6.3, we can exp<strong>and</strong> thelogarithm as a series <strong>and</strong> obtainlim ln y = lim[x(− ···)]2 a2x→∞ x→∞ x − a42 2x + = −a 2 .4There<strong>for</strong>e, since lim x→∞ ln y = −a 2 it follows that lim x→∞ y =exp(−a 2 ). ◭(v) L’Hôpital’s rule may be used; it is an extension of (iii)(c) above. In caseswhere both numerator <strong>and</strong> denominator are zero or both are infinite, furtherconsideration of the limit must follow. Let us first consider lim x→a f(x)/g(x),where f(a) =g(a) = 0. Exp<strong>and</strong>ing the numerator <strong>and</strong> denominator as Taylorseries we obtainf(x)g(x) = f(a)+(x − a)f′ (a)+[(x − a) 2 /2!]f ′′ g(a)+(x − a)g ′ (a)+[(x − a) 2 /2!]g ′′ (a)+···.However, f(a) =g(a) =0sof(x)g(x) = f′ (a)+[(x − a)/2!]f ′′ g ′ (a)+[(x − a)/2!]g ′′ (a)+···.142


4.7 EVALUATION OF LIMITSThere<strong>for</strong>e we findf(x)limx→a g(x) = f′ (a)g ′ (a) ,provided f ′ (a) <strong>and</strong> g ′ (a) are not themselves both equal to zero. If, however,f ′ (a) <strong>and</strong>g ′ (a) are both zero then the same process can be applied to the ratiof ′ (x)/g ′ (x) to yieldf(x)limx→a g(x) = f′′ (a)g ′′ (a) ,provided that at least one of f ′′ (a) <strong>and</strong>g ′′ (a) is non-zero. If the original limit doesexist then it can be found by repeating the process as many times as is necessary<strong>for</strong> the ratio of corresponding nth derivatives not to be of the indeterminate <strong>for</strong>m0/0, i.e.f(x)limx→a g(x) = f(n) (a)g (n) (a) .◮Evaluate the limitsin xlimx→0 x .We first note that if x = 0, both numerator <strong>and</strong> denominator are zero. Thus we applyl’Hôpital’s rule: differentiating, we obtainlim(sin x/x) = lim(cos x/1) = 1. ◭x→0 x→0So far we have only considered the case where f(a) =g(a) =0.Forthecasewhere f(a) =g(a) =∞ we may still apply l’Hôpital’s rule by writingf(x)limx→a g(x) = lim 1/g(x)x→a 1/f(x) ,whichisnowofthe<strong>for</strong>m0/0 atx = a. Note also that l’Hôpital’s rule is stillvalid <strong>for</strong> finding limits as x →∞,i.e.whena = ∞. This is easily shown by lettingy =1/x as follows:f(x)limx→∞ g(x) = lim f(1/y)y→0 g(1/y)−f ′ (1/y)/y 2= limy→0 −g ′ (1/y)/y 2f ′ (1/y)= limy→0 g ′ (1/y)f ′ (x)= limx→∞ g ′ (x) .143


SERIES AND LIMITSSummary of methods <strong>for</strong> evaluating limitsTo find the limit of a continuous function f(x) at a point x = a, simply substitute0the value a into the function noting that∞ = 0 <strong>and</strong> that ∞ 0= ∞. Theonlydifficulty occurs when either of the expressions 0 0 or ∞ ∞results. In this casedifferentiate top <strong>and</strong> bottom <strong>and</strong> try again. Continue differentiating until the top<strong>and</strong> bottom limits are no longer both zero or both infinity. If the undetermined<strong>for</strong>m 0 ×∞occurs then it can always be rewritten as 0 0 or ∞ ∞ .4.8 Exercises4.1 Sum the even numbers between 1000 <strong>and</strong> 2000 inclusive.4.2 If you invest £1000 on the first day of each year, <strong>and</strong> interest is paid at 5% onyour balance at the end of each year, how much money do you have after 25years?4.3 How does the convergence of the series∞∑ (n − r)!n!depend on the integer r?4.4 Show that <strong>for</strong> testing the convergence of the seriesn=rx + y + x 2 + y 2 + x 3 + y 3 + ··· ,where 0


4.8 EXERCISES4.8 The N + 1 complex numbers ω m are given by ω m =exp(2πim/N), <strong>for</strong> m =0, 1, 2,... ,N.(a) Evaluate the following:N∑N∑N∑(i) ω m , (ii) ωm,2 (iii) ω m x m .m=0m=0(b) Use these results to evaluate:N∑[ ( 2πm(i) cosNn=1)− cosm=0( )] 4πm, (ii)Nm=03∑( ) 2πm2 m sin .34.9 Prove thatcos θ +cos(θ + α)+···+cos(θ + nα) = sin 1 (n +1)α2sin 1 α cos(θ + 1 nα). 224.10 Determine whether the following series converge (θ <strong>and</strong> p are positive realnumbers):∞∑∞2sinnθ(a)n(n +1) , (b) ∑ 2n , (c) ∑ ∞12 2n , 1/2∞∑ (−1) n (n 2 +1) 1/2∞∑ n p(d), (e)n ln nn! .n=2n=14.11 Find the real values of x <strong>for</strong> which the following series are convergent:∞∑ x n∞(a)n +1 , (b) ∑∞∑(sin x) n , (c) n x ,n=1(d)∞∑e nx ,n=1n=1n=1(e)∞∑(ln n) x .4.12 Determine whether the following series are convergent:∞∑ n 1/2(a)(n +1) , (b) ∑ ∞n 21/2 n! , (c) ∑ ∞(ln n) n∞∑ n n, (d)n n/2 n! .n=1n=1n=1n=14.13 Determine whether the following series are absolutely convergent, convergent oroscillatory:∞∑ (−1) n(a)n , (b) ∑ ∞(−1) n (2n +1)∞∑ (−1) n |x| n, (c),5/2 nn!n=1n=1n=0∞∑ (−1) n(d)n 2 +3n +2 , (e) ∑ ∞(−1) n 2 n.n 1/2 n=0n=14.14 Obtain the positive values of x <strong>for</strong> which the following series converges:∞∑ x n/2 e −n.nn=1n=2m=0n=1n=1145


SERIES AND LIMITS4.15 Prove that∞∑n=2[ ]n r +(−1) nlnn ris absolutely convergent <strong>for</strong> r = 2, but only conditionally convergent <strong>for</strong> r =1.4.16 An extension to the proof of the integral test (subsection 4.3.2) shows that, if f(x)is positive, continuous <strong>and</strong> monotonically decreasing, <strong>for</strong> x ≥ 1, <strong>and</strong> the seriesf(1) + f(2) + ··· is convergent, then its sum does not exceed f(1) + L, whereLis the integral∫ ∞1f(x) dx.Use this result to show that the sum ζ(p) of the Riemann zeta series ∑ n −p ,withp>1, is not greater than p/(p − 1).4.17 Demonstrate that rearranging the order of its terms can make a conditionally∑ convergent series converge to a different limit by considering the series(−1) n+1 n −1 =ln2=0.693. Rearrange the series asS = 1 + 1 − 1 + 1 + 1 − 1 + 1 + 1 − 1 + 1 + ···1 3 2 5 7 4 9 11 6 13<strong>and</strong> group each set of three successive terms. Show that the series can then bewritten∞∑m=18m − 32m(4m − 3)(4m − 1) ,which is convergent (by comparison with ∑ n −2 ) <strong>and</strong> contains only positiveterms. Evaluate the first of these <strong>and</strong> hence deduce that S is not equal to ln 2.4.18 Illustrate result (iv) of section 4.4, concerning Cauchy products, by consideringthedoublesummationS =∞∑n∑n=1 r=11r 2 (n +1− r) 3 .By examining the points in the nr-plane over which the double summation is tobe carried out, show that S can be written asS =∞∑∞∑n=r r=11r 2 (n +1− r) 3 .Deduce that S ≤ 3.4.19 A Fabry–Pérot interferometer consists of two parallel heavily silvered glass plates;light enters normally to the plates, <strong>and</strong> undergoes repeated reflections betweenthem, with a small transmitted fraction emerging at each reflection. Find theintensity of the emerging wave, |B| 2 ,wherewith r <strong>and</strong> φ real.B = A(1 − r)146∞∑r n e inφ ,n=0


4.8 EXERCISES4.20 Identify the series∞∑ (−1) n+1 x 2n(2n − 1)! ,n=1<strong>and</strong> then, by integration <strong>and</strong> differentiation, deduce the values S of the followingseries:∞∑∞∑(a)(c)n=1∞∑n=1(−1) n+1 n 2, (b)(2n)!n=1(−1) n+1 nπ 2n∞4 n (2n − 1)! , (d) ∑n=0(−1) n+1 n(2n +1)! ,4.21 Starting from the Maclaurin series <strong>for</strong> cos x, show that(−1) n (n +1).(2n)!(cos x) −2 =1+x 2 + 2x43 + ··· .Deduce the first three terms in the Maclaurin series <strong>for</strong> tan x.4.22 Find the Maclaurin series <strong>for</strong>:( 1+x(a) ln1 − x4.23 Writing the nth derivative of f(x) =sinh −1 x as), (b) (x 2 +4) −1 , (c) sin 2 x.f (n) P n (x)(x) =(1 + x 2 ) , n−1/2where P n (x) is a polynomial (of order n − 1), show that the P n (x) satisfytherecurrence relationP n+1 (x) =(1+x 2 )P n(x) ′ − (2n − 1)xP n (x).Hence generate the coefficients necessary to express sinh −1 x as a Maclaurin seriesup to terms in x 5 .4.24 Find the first three non-zero terms in the Maclaurin series <strong>for</strong> the followingfunctions:(a) (x 2 +9) −1/2 , (b) ln[(2 + x) 3 ], (c) exp(sin x),(d) ln(cos x), (e) exp[−(x − a) −2 ], (f) tan −1 x.4.25 By using the logarithmic series, prove that if a <strong>and</strong> b are positive <strong>and</strong> nearlyequal thenln a 2(a − b)≃b a + b .Show that the error in this approximation is about 2(a − b) 3 /[3(a + b) 3 ].4.26 Determine whether the following functions f(x) are (i) continuous, <strong>and</strong> (ii)differentiable at x =0:(a) f(x) =exp(−|x|);(b) f(x) =(1− cos x)/x 2 <strong>for</strong> x ≠0,f(0) = 1 ; 2(c) f(x) =x sin(1/x) <strong>for</strong>x ≠0,f(0) = 0;(d) f(x) =[4− x 2 ], where [y] denotes the integer part of y.4.27 Find the limit as x → 0of[ √ 1+x m − √ 1 − x m ]/x n ,inwhichm <strong>and</strong> n are positiveintegers.4.28 Evaluate the following limits:147


SERIES AND LIMITSsin 3x(a) limx→0 sinh x ,tan x − x(c) limx→0 cos x − 1 ,tan x − tanh x(b) limx→0 sinh x − x ,( cosec x(d) lim − sinh x ).x→0 x 3 x 54.29 Find the limits of the following functions:x 3 + x 2 − 5x − 2(a), as x → 0, x →∞<strong>and</strong> x → 2;2x 3 − 7x 2 +4x +4sin x − x cosh x(b), as x → 0;sinh x − x∫ π/2( )y cos y − sin y(c)dy, as x → 0.x y 24.30 Use Taylor expansions to three terms to find approximations to (a) 4√ 17, <strong>and</strong>(b) 3√ 26.4.31 Using a first-order Taylor expansion about x = x 0 , show that a better approximationthan x 0 to the solution of the equationf(x) =sinx +tanx =2is given by x = x 0 + δ, where2 − f(x 0 )δ =.cos x 0 +sec 2 x 0(a) Use this procedure twice to find the solution of f(x) = 2 to six significantfigures, given that it is close to x =0.9.(b) Use the result in (a) to deduce, to the same degree of accuracy, one solutionof the quartic equationy 4 − 4y 3 +4y 2 +4y − 4=0.4.32 Evaluate[ ( 1lim cosec x − 1x→0 x 3 x − x )].64.33 In quantum theory, a system of oscillators, each of fundamental frequency ν <strong>and</strong>interacting at temperature T , has an average energy Ē given by∑ ∞n=0Ē = ∑ nhνe−nx∞ ,n=0 e−nxwhere x = hν/kT , h <strong>and</strong> k being the Planck <strong>and</strong> Boltzmann constants, respectively.Prove that both series converge, evaluate their sums, <strong>and</strong> show that at hightemperatures Ē ≈ kT, whilst at low temperatures Ē ≈ hν exp(−hν/kT ).4.34 In a very simple model of a crystal, point-like atomic ions are regularly spacedalong an infinite one-dimensional row with spacing R. Alternate ions carry equal<strong>and</strong> opposite charges ±e. The potential energy of the ith ion in the electric fielddue to another ion, the jth, isq i q j,4πɛ 0 r ijwhere q i , q j are the charges on the ions <strong>and</strong> r ij is the distance between them.Write down a series giving the total contribution V i of the ithiontotheoverallpotential energy. Show that the series converges, <strong>and</strong>, if V i is written asV i =148αe24πɛ 0 R ,


4.9 HINTS AND ANSWERSfind a closed-<strong>for</strong>m expression <strong>for</strong> α, the Madelung constant <strong>for</strong> this (unrealistic)lattice.4.35 One of the factors contributing to the high relative permittivity of water to staticelectric fields is the permanent electric dipole moment, p, of the water molecule.In an external field E the dipoles tend to line up with the field, but they do notdo so completely because of thermal agitation corresponding to the temperature,T , of the water. A classical (non-quantum) calculation using the Boltzmanndistribution shows that the average polarisability per molecule, α, isgivenbyα = p E (coth x − x−1 ),where x = pE/(kT) <strong>and</strong>k is the Boltzmann constant.At ordinary temperatures, even with high field strengths (10 4 Vm −1 or more),x ≪ 1. By making suitable series expansions of the hyperbolic functions involved,show that α = p 2 /(3kT) to an accuracy of about one part in 15x −2 .4.36 In quantum theory, a certain method (the Born approximation) gives the (socalled)amplitude f(θ) <strong>for</strong> the scattering of a particle of mass m through an angleθ by a uni<strong>for</strong>m potential well of depth V 0 <strong>and</strong> radius b (i.e. the potential energyof the particle is −V 0 within a sphere of radius b <strong>and</strong> zero elsewhere) asf(θ) = 2mV 0(sin Kb − Kbcos Kb). 2 K3 Here is the Planck constant divided by 2π, the energy of the particle is 2 k 2 /(2m)<strong>and</strong> K is 2k sin(θ/2).Use l’Hôpital’s rule to evaluate the amplitude at low energies, i.e. when k <strong>and</strong>hence K tend to zero, <strong>and</strong> so determine the low-energy total cross-section.[ Note: the differential cross-section is given by |f(θ)| 2 <strong>and</strong> the total crosssectionby the integral of this over all solid angles, i.e. 2π ∫ π0 |f(θ)|2 sin θdθ.]4.9 Hints <strong>and</strong> answers4.1 Write as 2( ∑ 1000n=1 n − ∑ 499n=1n) = 751500.4.3 Divergent <strong>for</strong> r ≤ 1; convergent <strong>for</strong> r ≥ 2.4.5 (a) ln(N + 1), divergent; (b) 1 [1 − 3 (−2)n ], oscillates infinitely; (c) Add 1 S 3 N to theS N series; 3 [1 − 16 (−3)−N ]+ 3 4 N(−3)−N−1 ,convergentto 3 . 164.7 Write the nth term as the difference between two consecutive values of a partialfractionfunction of n. Thesumequals 1 (1 − 2 N−2 ).4.9 Sum the geometric series with rth term exp[i(θ + rα)]. Its real part is{cos θ − cos [(n +1)α + θ] − cos(θ − α)+cos(θ + nα)} /4sin 2 (α/2),which can be reduced to the given answer.4.11 (a) −1 ≤ x


SERIES AND LIMITS4.15 Divide the series into two series, n odd <strong>and</strong> n even. For r = 2 both are absolutelyconvergent, by comparison with ∑ n −2 .Forr = 1 neither series is convergent,by comparison with ∑ n −1 . However, the sum of the two is convergent, by thealternating sign test or by showing that the terms cancel in pairs.4.17 The first term has value 0.833 <strong>and</strong> all other terms are positive.4.19 |A| 2 (1 − r) 2 /(1 + r 2 − 2r cos φ).4.21 Use the binomial expansion <strong>and</strong> collect terms up to x 4 . Integrate both sides ofthe displayed equation. tan x = x + x 3 /3+2x 5 /15 + ···.4.23 For example, P 5 (x) =24x 4 − 72x 2 +9.sinh −1 x = x − x 3 /6+3x 5 /40 − ···.4.25 Set a = D + δ <strong>and</strong> b = D − δ <strong>and</strong> use the expansion <strong>for</strong> ln(1 ± δ/D).4.27 The limit is 0 <strong>for</strong> m>n,1<strong>for</strong>m = n, <strong>and</strong>∞ <strong>for</strong> m


5Partial differentiationIn chapter 2, we discussed functions f of only one variable x, which were usuallywritten f(x). Certain constants <strong>and</strong> parameters may also have appeared in thedefinition of f, e.g.f(x) =ax+2 contains the constant 2 <strong>and</strong> the parameter a, butonly x was considered as a variable <strong>and</strong> only the derivatives f (n) (x) =d n f/dx nwere defined.However, we may equally well consider functions that depend on more than onevariable, e.g. the function f(x, y) =x 2 +3xy, which depends on the two variablesx <strong>and</strong> y. For any pair of values x, y, the function f(x, y) has a well-defined value,e.g. f(2, 3) = 22. This notion can clearly be extended to functions dependent onmore than two variables. For the n-variable case, we write f(x 1 ,x 2 ,...,x n )<strong>for</strong>a function that depends on the variables x 1 ,x 2 ,...,x n .Whenn =2,x 1 <strong>and</strong> x 2correspond to the variables x <strong>and</strong> y used above.Functions of one variable, like f(x), can be represented by a graph on aplane sheet of paper, <strong>and</strong> it is apparent that functions of two variables can,with little ef<strong>for</strong>t, be represented by a surface in three-dimensional space. Thus,we may also picture f(x, y) as describing the variation of height with positionin a mountainous l<strong>and</strong>scape. Functions of many variables, however, are usuallyvery difficult to visualise <strong>and</strong> so the preliminary discussion in this chapter willconcentrate on functions of just two variables.5.1 Definition of the partial derivativeIt is clear that a function f(x, y) of two variables will have a gradient in alldirections in the xy-plane. A general expression <strong>for</strong> this rate of change can befound <strong>and</strong> will be discussed in the next section. However, we first consider thesimpler case of finding the rate of change of f(x, y) in the positive x- <strong>and</strong>ydirections.These rates of change are called the partial derivatives with respect151


PARTIAL DIFFERENTIATIONto x <strong>and</strong> y respectively, <strong>and</strong> they are extremely important in a wide range ofphysical applications.For a function of two variables f(x, y) we may define the derivative with respectto x, <strong>for</strong> example, by saying that it is that <strong>for</strong> a one-variable function when y isheld fixed <strong>and</strong> treated as a constant. To signify that a derivative is with respectto x, but at the same time to recognize that a derivative with respect to y alsoexists, the <strong>for</strong>mer is denoted by ∂f/∂x <strong>and</strong> is the partial derivative of f(x, y) withrespect to x. Similarly, the partial derivative of f with respect to y is denoted by∂f/∂y.To define <strong>for</strong>mally the partial derivative of f(x, y) with respect to x, we have∂f∂x = lim∆x→0f(x +∆x, y) − f(x, y), (5.1)∆xprovided that the limit exists. This is much the same as <strong>for</strong> the derivative of aone-variable function. The other partial derivative of f(x, y) is similarly definedas a limit (provided it exists):∂f∂y = lim∆y→0f(x, y +∆y) − f(x, y). (5.2)∆yIt is common practice in connection with partial derivatives of functionsinvolving more than one variable to indicate those variables that are held constantby writing them as subscripts to the derivative symbol. Thus, the partial derivativesdefined in (5.1) <strong>and</strong> (5.2) would be written respectively as( ∂f∂x)y<strong>and</strong>( ∂f∂yIn this <strong>for</strong>m, the subscript shows explicitly which variable is to be kept constant.A more compact notation <strong>for</strong> these partial derivatives is f x <strong>and</strong> f y . However, it isextremely important when using partial derivatives to remember which variablesare being held constant <strong>and</strong> it is wise to write out the partial derivative in explicit<strong>for</strong>m if there is any possibility of confusion.The extension of the definitions (5.1), (5.2) to the general n-variable case isstraight<strong>for</strong>ward <strong>and</strong> can be written <strong>for</strong>mally as∂f(x 1 ,x 2 ,...,x n )∂x i[f(x 1 ,x 2 ,...,x i +∆x i ,...,x n ) − f(x 1 ,x 2 ,...,x i ,...,x n )]= lim,∆x i→0∆x iprovided that the limit exists.Just as <strong>for</strong> one-variable functions, second (<strong>and</strong> higher) partial derivatives maybe defined in a similar way. For a two-variable function f(x, y) theyare(∂ ∂f∂x ∂x(∂ ∂f∂x ∂y)= ∂2 f∂x 2 = f xx,)= ∂2 f∂x∂y = f xy,152∂∂y∂∂y( ∂f∂y( ∂f∂x).x)= ∂2 f∂y 2 = f yy,)= ∂2 f∂y∂x = f yx.


5.2 THE TOTAL DIFFERENTIAL AND TOTAL DERIVATIVEOnly three of the second derivatives are independent since the relation∂ 2 f∂x∂y = ∂2 f∂y∂x ,is always obeyed, provided that the second partial derivatives are continuousat the point in question. This relation often proves useful as a labour-savingdevice when evaluating second partial derivatives. It can also be shown that <strong>for</strong>a function of n variables, f(x 1 ,x 2 ,...,x n ), under the same conditions,∂ 2 f=∂2 f.∂x i ∂x j ∂x j ∂x i◮Find the first <strong>and</strong> second partial derivatives of the functionf(x, y) =2x 3 y 2 + y 3 .The first partial derivatives are∂f∂x =6x2 y 2 ,<strong>and</strong> the second partial derivatives are∂ 2 f∂x 2 =12xy2 ,∂ 2 f∂y 2 =4x3 +6y,the last two being equal, as expected. ◭∂f∂y =4x3 y +3y 2 ,∂ 2 f∂x∂y =12x2 y,∂ 2 f∂y∂x =12x2 y,5.2 The total differential <strong>and</strong> total derivativeHaving defined the (first) partial derivatives of a function f(x, y), which give therate of change of f along the positive x- <strong>and</strong>y-axes, we consider next the rate ofchange of f(x, y) in an arbitrary direction. Suppose that we make simultaneoussmall changes ∆x in x <strong>and</strong> ∆y in y <strong>and</strong> that, as a result, f changes to f +∆f.Then we must have∆f = f(x +∆x, y +∆y) − f(x, y)= f(x +∆x, y +∆y) − f(x, y +∆y)+f(x, y +∆y) − f(x, y)[ ] [ ]f(x +∆x, y +∆y) − f(x, y +∆y) f(x, y +∆y) − f(x, y)=∆x +∆y.∆x∆y(5.3)In the last line we note that the quantities in brackets are very similar to thoseinvolved in the definitions of partial derivatives (5.1), (5.2). For them to be strictlyequal to the partial derivatives, ∆x <strong>and</strong> ∆y would need to be infinitesimally small.But even <strong>for</strong> finite (but not too large) ∆x <strong>and</strong> ∆y the approximate <strong>for</strong>mula∆f ≈∂f(x, y)∆x +∂x153∂f(x, y)∆y, (5.4)∂y


PARTIAL DIFFERENTIATIONcan be obtained. It will be noticed that the first bracket in (5.3) actually approximatesto ∂f(x, y +∆y)/∂x but that this has been replaced by ∂f(x, y)/∂x in (5.4).This approximation clearly has the same degree of validity as that which replacesthe bracket by the partial derivative.How valid an approximation (5.4) is to (5.3) depends not only on how small∆x <strong>and</strong> ∆y are but also on the magnitudes of higher partial derivatives; this isdiscussed further in section 5.7 in the context of Taylor series <strong>for</strong> functions ofmore than one variable. Nevertheless, letting the small changes ∆x <strong>and</strong> ∆y in(5.4) become infinitesimal, we can define the total differential df of the functionf(x, y), without any approximation, asdf = ∂f ∂fdx + dy. (5.5)∂x ∂yEquation (5.5) can be extended to the case of a function of n variables,f(x 1 ,x 2 ,...,x n );df = ∂f dx 1 + ∂f dx 2 + ···+ ∂f dx n . (5.6)∂x 1 ∂x 2 ∂x n◮Find the total differential of the function f(x, y) =y exp(x + y).Evaluating the first partial derivatives, we find∂f∂f= y exp(x + y),∂x=exp(x + y)+y exp(x + y).∂yApplying (5.5), we then find that the total differential is given bydf =[y exp(x + y)]dx +[(1+y)exp(x + y)]dy. ◭In some situations, despite the fact that several variables x i , i =1, 2,...,n,appear to be involved, effectively only one of them is. This occurs if there aresubsidiary relationships constraining all the x i to have values dependent on thevalue of one of them, say x 1 . These relationships may be represented by equationsthat are typically of the <strong>for</strong>mx i = x i (x 1 ), i =2, 3,...,n. (5.7)In principle f can then be expressed as a function of x 1 alone by substitutingfrom (5.7) <strong>for</strong> x 2 ,x 3 ,...,x n , <strong>and</strong> then the total derivative (or simply the derivative)of f with respect to x 1 is obtained by ordinary differentiation.Alternatively, (5.6) can be used to givedfdx 1= ∂f∂x 1+( ∂f∂x 2) dx2dx 1+ ···+( ∂f∂x n) dxndx 1. (5.8)It should be noted that the LHS of this equation is the total derivative df/dx 1 ,whilst the partial derivative ∂f/∂x 1 <strong>for</strong>ms only a part of the RHS. In evaluating154


5.3 EXACT AND INEXACT DIFFERENTIALSthis partial derivative account must be taken only of explicit appearances of x 1 inthe function f, <strong>and</strong>no allowance must be made <strong>for</strong> the knowledge that changingx 1 necessarily changes x 2 ,x 3 ,...,x n . The contribution from these latter changes isprecisely that of the remaining terms on the RHS of (5.8). Naturally, what hasbeen shown using x 1 in the above argument applies equally well to any other ofthe x i , with the appropriate consequent changes.◮Find the total derivative of f(x, y) =x 2 +3xy with respect to x, given that y =sin −1 x.We can see immediately that∂f∂f=2x +3y,∂x∂y =3x,dydx = 1(1 − x 2 ) 1/2<strong>and</strong> so, using (5.8) with x 1 = x <strong>and</strong> x 2 = y,dfdx =2x +3y +3x 1(1 − x 2 ) 1/2=2x +3sin −1 3xx +(1 − x 2 ) . 1/2Obviously the same expression would have resulted if we had substituted <strong>for</strong> y from thestart, but the above method often produces results with reduced calculation, particularlyin more complicated examples. ◭5.3 Exact <strong>and</strong> inexact differentialsIn the last section we discussed how to find the total differential of a function, i.e.its infinitesimal change in an arbitrary direction, in terms of its gradients ∂f/∂x<strong>and</strong> ∂f/∂y in the x- <strong>and</strong>y- directions (see (5.5)). Sometimes, however, we wishto reverse the process <strong>and</strong> find the function f that differentiates to give a knowndifferential. Usually, finding such functions relies on inspection <strong>and</strong> experience.As an example, it is easy to see that the function whose differential is df =xdy+ ydx is simply f(x, y) =xy + c, wherec is a constant. Differentials such asthis, which integrate directly, are called exact differentials, whereas those that donot are inexact differentials. For example, xdy+3ydx is not the straight<strong>for</strong>warddifferential of any function (see below). Inexact differentials can be made exact,however, by multiplying through by a suitable function called an integratingfactor. This is discussed further in subsection 14.2.3.◮Show that the differential xdy+3ydx is inexact.On the one h<strong>and</strong>, if we integrate with respect to x we conclude that f(x, y) =3xy + g(y),where g(y) is any function of y. On the other h<strong>and</strong>, if we integrate with respect to y weconclude that f(x, y) =xy + h(x) whereh(x) is any function of x. These conclusions areinconsistent <strong>for</strong> any <strong>and</strong> every choice of g(y) <strong>and</strong>h(x), <strong>and</strong> there<strong>for</strong>e the differential isinexact. ◭It is naturally of interest to investigate which properties of a differential make155


PARTIAL DIFFERENTIATIONit exact. Consider the general differential containing two variables,df = A(x, y) dx + B(x, y) dy.We see that∂f∂f= A(x, y),∂x= B(x, y)∂y<strong>and</strong>, using the property f xy = f yx , we there<strong>for</strong>e require∂A∂y = ∂B∂x . (5.9)This is in fact both a necessary <strong>and</strong> a sufficient condition <strong>for</strong> the differential tobe exact.◮Using (5.9) show that xdy+3ydx is inexact.In the above notation, A(x, y) =3y <strong>and</strong> B(x, y) =x <strong>and</strong> so∂A∂y =3,∂B∂x =1.As these are not equal it follows that the differential is inexact. ◭Determining whether a differential containing many variable x 1 ,x 2 ,...,x n isexact is a simple extension of the above. A differential containing many variablescanbewritteningeneralas<strong>and</strong> will be exact ifdf =n∑g i (x 1 ,x 2 ,...,x n ) dx ii=1∂g i∂x j= ∂g j∂x i<strong>for</strong> all pairs i, j. (5.10)There will be 1 2n(n − 1) such relationships to be satisfied.◮Show thatis an exact differential.(y + z) dx + xdy+ xdzIn this case, g 1 (x, y, z) =y + z, g 2 (x, y, z) =x, g 3 (x, y, z) =x <strong>and</strong> hence ∂g 1 /∂y =1=∂g 2 /∂x, ∂g 3 /∂x =1=∂g 1 /∂z, ∂g 2 /∂z =0=∂g 3 /∂y; there<strong>for</strong>e, from (5.10), the differentialis exact. As mentioned above, it is sometimes possible to show that a differential is exactsimply by finding by inspection the function from which it originates. In this example, itcan be seen easily that f(x, y, z) =x(y + z)+c. ◭156


5.4 USEFUL THEOREMS OF PARTIAL DIFFERENTIATION5.4 Useful theorems of partial differentiationSo far our discussion has centred on a function f(x, y) dependent on two variables,x <strong>and</strong> y. Equally, however, we could have expressed x as a function of f <strong>and</strong> y,or y as a function of f <strong>and</strong> x. To emphasise the point that all the variables areof equal st<strong>and</strong>ing, we now replace f by z. This does not imply that x, y <strong>and</strong> zare coordinate positions (though they might be). Since x is a function of y <strong>and</strong> z,it follows thatdx =( ) ∂xdy +∂yz( ) ∂xdz (5.11)∂zy<strong>and</strong> similarly, since y = y(x, z),( ) ( )∂y ∂ydy = dx + dz. (5.12)∂xz∂zxWe may now substitute (5.12) into (5.11) to obtain( ) ( ) [ (∂x ) ( ) ( ) ]∂x ∂y∂y ∂xdx =dx ++ dz. (5.13)∂yz∂xz∂yz∂zx∂zyNow if we hold z constant, so that dz = 0, we obtain the reciprocity relation( ) ( ) −1∂x ∂y= ,∂yz∂xzwhich holds provided both partial derivatives exist <strong>and</strong> neither is equal to zero.Note, further, that this relationship only holds when the variable being keptconstant, in this case z, is the same on both sides of the equation.Alternatively we can put dx = 0 in (5.13). Then the contents of the squarebrackets also equal zero, <strong>and</strong> we obtain the cyclic relation( ) ( ) ( )∂y ∂z ∂x= −1,∂zx∂xy∂yzwhich holds unless any of the derivatives vanish. In deriving this result we haveused the reciprocity relation to replace (∂x/∂z) −1y by (∂z/∂x) y .5.5 The chain ruleSo far we have discussed the differentiation of a function f(x, y) with respect toits variables x <strong>and</strong> y. We now consider the case where x <strong>and</strong> y are themselvesfunctions of another variable, say u. If we wish to find the derivative df/du,we could simply substitute in f(x, y) the expressions <strong>for</strong> x(u) <strong>and</strong>y(u) <strong>and</strong>thendifferentiate the resulting function of u. Such substitution will quickly give thedesired answer in simple cases, but in more complicated examples it is easier tomake use of the total differentials described in the previous section.157


PARTIAL DIFFERENTIATIONFrom equation (5.5) the total differential of f(x, y) is given bydf = ∂f ∂fdx +∂x ∂y dy,but we now note that by using the <strong>for</strong>mal device of dividing through by du thisimmediately impliesdfdu = ∂f dx∂x du + ∂f dy∂y du , (5.14)which is called the chain rule <strong>for</strong> partial differentiation. This expression providesa direct method <strong>for</strong> calculating the total derivative of f with respect to u <strong>and</strong> isparticularly useful when an equation is expressed in a parametric <strong>for</strong>m.◮Given that x(u) =1+au <strong>and</strong> y(u) =bu 3 , find the rate of change of f(x, y) =xe −y withrespect to u.As discussed above, this problem could be addressed by substituting <strong>for</strong> x <strong>and</strong> y to obtainf as a function only of u <strong>and</strong> then differentiating with respect to u. However, using (5.14)directly we obtaindfdu =(e−y )a +(−xe −y )3bu 2 ,which on substituting <strong>for</strong> x <strong>and</strong> y givesdfdu = e−bu3 (a − 3bu 2 − 3bau 3 ). ◭Equation (5.14) is an example of the chain rule <strong>for</strong> a function of two variableseach of which depends on a single variable. The chain rule may be extended tofunctions of many variables, each of which is itself a function of a variable u, i.e.f(x 1 ,x 2 ,x 3 ,...,x n ), with x i = x i (u). In this case the chain rule givesdfdu =n∑i=1∂f dx i∂x i du = ∂f dx 1∂x 1 du + ∂f dx 2 ∂f dx n+ ···+∂x 2 du ∂x n du . (5.15)5.6 Change of variablesIt is sometimes necessary or desirable to make a change of variables during thecourse of an analysis, <strong>and</strong> consequently to have to change an equation expressedin one set of variables into an equation using another set. The same situation arisesif a function f depends on one set of variables x i ,sothatf = f(x 1 ,x 2 ,...,x n ) butthe x i are themselves functions of a further set of variables u j <strong>and</strong> given by theequationsx i = x i (u 1 ,u 2 ,...,u m ). (5.16)158


5.6 CHANGE OF VARIABLESyρφxFigure 5.1The relationship between Cartesian <strong>and</strong> plane polar coordinates.For each different value of i, x i will be a different function of the u j .Inthiscasethe chain rule (5.15) becomes∂f∂u j=n∑i=1∂f∂x i∂x i∂u j, j =1, 2,...,m, (5.17)<strong>and</strong> is said to express a change of variables. In general the number of variablesin each set need not be equal, i.e. m need not equal n, but if both the x i <strong>and</strong> theu i are sets of independent variables then m = n.◮Plane polar coordinates, ρ <strong>and</strong> φ, <strong>and</strong> Cartesian coordinates, x <strong>and</strong> y, arerelatedbytheexpressionsx = ρ cos φ, y = ρ sin φ,as can be seen from figure 5.1. An arbitrary function f(x, y) canbere-expressedasafunction g(ρ, φ). Trans<strong>for</strong>m the expression∂ 2 f∂x + ∂2 f2 ∂y 2into one in ρ <strong>and</strong> φ.We first note that ρ 2 = x 2 + y 2 , φ =tan −1 (y/x). We can now write down the four partialderivatives∂ρ∂x = x(x 2 + y 2 ) =cosφ, ∂φ1/2 ∂x = −(y/x2 )1+(y/x) = − sin φ2 ρ ,∂ρ∂y = y(x 2 + y 2 ) =sinφ, ∂φ1/2 ∂y = 1/x1+(y/x) = cos φ2 ρ .159


PARTIAL DIFFERENTIATIONThus, from (5.17), we may write∂∂x =cosφ ∂∂ρ − sin φ ∂ρ ∂φ ,Now it is only a matter of writing∂ 2 f∂x = ∂ ( ) ∂f= ∂ ( ∂2 ∂x ∂x ∂x ∂x(= cos φ ∂∂ρ − sin φ ∂ρ ∂φ(= cos φ ∂ ∂∂ρ − sin φρ ∂φ)f∂∂y =sinφ ∂∂ρ + cos φρ)(cos φ ∂)(cos φ ∂g=cos 2 φ ∂2 g 2cosφ sin φ ∂g+∂ρ2 ρ 2 ∂φ+ sin2 φ ∂gρ ∂ρ + sin2 φ ∂ 2 gρ 2 ∂φ 2∂ρ − sin φ ∂ρ ∂φ∂ρ − sin φ )∂gρ ∂φ−2cosφ sin φρ<strong>and</strong> a similar expression <strong>for</strong> ∂ 2 f/∂y 2 ,(∂ 2 f∂y = sin φ ∂2 ∂ρ + cos φ )(∂sin φ ∂ρ ∂φ ∂ρ + cos φρ=sin 2 φ ∂2 g 2cosφ sin φ ∂g−∂ρ2 ρ 2 ∂φ+ cos2 φ ∂gρ ∂ρ + cos2 φ ∂ 2 gρ 2 ∂φ . 2+2cosφ sin φρ∂∂φ .)g∂ 2 g∂φ∂ρ∂∂φ)g∂ 2 g∂φ∂ρWhen these two expressions are added together the change of variables is complete <strong>and</strong>we obtain∂ 2 f∂x + ∂2 f2 ∂y = ∂2 g2 ∂ρ + 1 ∂g2 ρ ∂ρ + 1 ∂ 2 gρ 2 ∂φ . ◭ 25.7 Taylor’s theorem <strong>for</strong> many-variable functionsWe have already introduced Taylor’s theorem <strong>for</strong> a function f(x) of one variable,in section 4.6. In an analogous way, the Taylor expansion of a function f(x, y) oftwo variables is given byf(x, y) =f(x 0 ,y 0 )+ ∂f ∂f∆x +∂x ∂y ∆y+ 1 [ ∂ 2 ]f2! ∂x 2 (∆x)2 +2 ∂2 f∂x∂y ∆x∆y + ∂2 f∂y 2 (∆y)2 + ··· , (5.18)where ∆x = x − x 0 <strong>and</strong> ∆y = y − y 0 , <strong>and</strong> all the derivatives are to be evaluatedat (x 0 ,y 0 ).160


5.7 TAYLOR’S THEOREM FOR MANY-VARIABLE FUNCTIONS◮Find the Taylor expansion, up to quadratic terms in x − 2 <strong>and</strong> y − 3, off(x, y) =y exp xyabout the point x =2, y =3.We first evaluate the required partial derivatives of the function, i.e.∂f∂x = y2 exp xy,∂f=expxy + xy exp xy,∂y∂ 2 f∂x 2 = y3 exp xy,∂ 2 f∂y 2 =2x exp xy + x2 y exp xy,∂ 2 f∂x∂y =2y exp xy + xy2 exp xy.Using (5.18), the Taylor expansion of a two-variable function, we findf(x, y) ≈ e 6 {3+9(x − 2) + 7(y − 3)+(2!) −1 [ 27(x − 2) 2 + 48(x − 2)(y − 3) + 16(y − 3) 2]} . ◭It will be noticed that the terms in (5.18) containing first derivatives can bewritten as(∂f ∂f∆x +∂x ∂y ∆y = ∆x ∂∂x +∆y ∂ )f(x, y),∂ywhere both sides of this relation should be evaluated at the point (x 0 ,y 0 ). Similarlythe terms in (5.18) containing second derivatives can be written as[1 ∂ 2 ]f2! ∂x 2 (∆x)2 +2 ∂2 f∂x∂y ∆x∆y + ∂2 f∂y 2 (∆y)2 = 1 (∆x ∂2! ∂x +∆y ∂ ) 2f(x, y),∂y(5.19)where it is understood that the partial derivatives resulting from squaring theexpression in parentheses act only on f(x, y) <strong>and</strong> its derivatives, <strong>and</strong> not on ∆xor ∆y; again both sides of (5.19) should be evaluated at (x 0 ,y 0 ). It can be shownthat the higher-order terms of the Taylor expansion of f(x, y) can be written inan analogous way, <strong>and</strong> that we may write the full Taylor series asf(x, y) =∞∑n=0[(1∆x ∂n! ∂x +∆y ∂ ) nf(x, y)]∂yx 0,y 0where, as indicated, all the terms on the RHS are to be evaluated at (x 0 ,y 0 ).The most general <strong>for</strong>m of Taylor’s theorem, <strong>for</strong> a function f(x 1 ,x 2 ,...,x n )ofnvariables, is a simple extension of the above. Although it is not necessary to doso, we may think of the x i as coordinates in n-dimensional space <strong>and</strong> write thefunction as f(x), where x is a vector from the origin to (x 1 ,x 2 ,...,x n ). Taylor’s161


PARTIAL DIFFERENTIATIONtheorem then becomesf(x) =f(x 0 )+ ∑ i∂f∂x i∆x i + 1 2!∑ ∑ij∂ 2 f∂x i ∂x j∆x i ∆x j + ··· ,(5.20)where ∆x i = x i − x i0 <strong>and</strong> the partial derivatives are evaluated at (x 10 ,x 20 ,...,x n0 ).For completeness, we note that in this case the full Taylor series can be writtenin the <strong>for</strong>mf(x) =∞∑n=01 [(∆x · ∇) n f(x) ] n!x=x 0,where ∇ is the vector differential operator del, to be discussed in chapter 10.5.8 Stationary values of many-variable functionsThe idea of the stationary points of a function of just one variable has alreadybeen discussed in subsection 2.1.8. We recall that the function f(x) has a stationarypoint at x = x 0 if its gradient df/dx is zero at that point. A function may haveany number of stationary points, <strong>and</strong> their nature, i.e. whether they are maxima,minima or stationary points of inflection, is determined by the value of the secondderivative at the point. A stationary point is(i) a minimum if d 2 f/dx 2 > 0;(ii) a maximum if d 2 f/dx 2 < 0;(iii) a stationary point of inflection if d 2 f/dx 2 = 0 <strong>and</strong> changes sign throughthe point.We now consider the stationary points of functions of more than one variable;we will see that partial differential analysis is ideally suited to the determinationof the position <strong>and</strong> nature of such points. It is helpful to consider first the caseof a function of just two variables but, even in this case, the general situationis more complex than that <strong>for</strong> a function of one variable, as can be seen fromfigure 5.2.This figure shows part of a three-dimensional model of a function f(x, y). Atpositions P <strong>and</strong> B there are a peak <strong>and</strong> a bowl respectively or, more mathematically,a local maximum <strong>and</strong> a local minimum. At position S the gradient in anydirection is zero but the situation is complicated, since a section parallel to theplane x = 0 would show a maximum, but one parallel to the plane y =0wouldshow a minimum. A point such as S is known as a saddle point. The orientationof the ‘saddle’ in the xy-plane is irrelevant; it is as shown in the figure solely <strong>for</strong>ease of discussion. For any saddle point the function increases in some directionsaway from the point but decreases in other directions.162


5.8 STATIONARY VALUES OF MANY-VARIABLE FUNCTIONSPSyxBFigure 5.2 Stationary points of a function of two variables. A minimumoccurs at B, a maximum at P <strong>and</strong> a saddle point at S.For functions of two variables, such as the one shown, it should be clear that anecessary condition <strong>for</strong> a stationary point (maximum, minimum or saddle point)to occur is that∂f∂x = 0 <strong>and</strong> ∂f=0. (5.21)∂yThe vanishing of the partial derivatives in directions parallel to the axes is enoughto ensure that the partial derivative in any arbitrary direction is also zero. Thelatter can be considered as the superposition of two contributions, one alongeach axis; since both contributions are zero, so is the partial derivative in thearbitrary direction. This may be made more precise by considering the totaldifferentialdf = ∂f ∂fdx +∂x ∂y dy.Using (5.21) we see that although the infinitesimal changes dx <strong>and</strong> dy can bechosen independently the change in the value of the infinitesimal function df isalways zero at a stationary point.We now turn our attention to determining the nature of a stationary point ofa function of two variables, i.e. whether it is a maximum, a minimum or a saddlepoint. By analogy with the one-variable case we see that ∂ 2 f/∂x 2 <strong>and</strong> ∂ 2 f/∂y 2must both be positive <strong>for</strong> a minimum <strong>and</strong> both be negative <strong>for</strong> a maximum.However these are not sufficient conditions since they could also be obeyed atcomplicated saddle points. What is important <strong>for</strong> a minimum (or maximum) isthat the second partial derivative must be positive (or negative) in all directions,not just in the x- <strong>and</strong>y- directions.163


PARTIAL DIFFERENTIATIONTo establish just what constitutes sufficient conditions we first note that, sincef is a function of two variables <strong>and</strong> ∂f/∂x = ∂f/∂y = 0, a Taylor expansion ofthe type (5.18) about the stationary point yieldsf(x, y) − f(x 0 ,y 0 ) ≈ 1 2![(∆x) 2 f xx +2∆x∆yf xy +(∆y) 2 f yy],where ∆x = x − x 0 <strong>and</strong> ∆y = y − y 0 <strong>and</strong> where the partial derivatives have beenwritten in more compact notation. Rearranging the contents of the bracket asthe weighted sum of two squares, we findf(x, y) − f(x 0 ,y 0 ) ≈ 1 2( [f xx ∆x + f ) ( )]2xy∆y+(∆y) 2 f yy − f2 xy.f xx f xx(5.22)For a minimum, we require (5.22) to be positive <strong>for</strong> all ∆x <strong>and</strong> ∆y, <strong>and</strong> hencef xx > 0<strong>and</strong>f yy − (fxy/f 2 xx ) > 0. Given the first constraint, the second can bewritten f xx f yy >fxy. 2 Similarly <strong>for</strong> a maximum we require (5.22) to be negative,<strong>and</strong> hence f xx < 0<strong>and</strong>f xx f yy >fxy 2 . For minima <strong>and</strong> maxima, symmetry requiresthat f yy obeys the same criteria as f xx . When (5.22) is negative (or zero) <strong>for</strong> somevalues of ∆x <strong>and</strong> ∆y but positive (or zero) <strong>for</strong> others, we have a saddle point. Inthis case f xx f yy


5.8 STATIONARY VALUES OF MANY-VARIABLE FUNCTIONS◮Show that the function f(x, y) =x 3 exp(−x 2 −y 2 ) has a maximum at the point ( √ 3/2, 0),a minimum at (− √ 3/2, 0) <strong>and</strong> a stationary point at the origin whose nature cannot bedetermined by the above procedures.Setting the first two partial derivatives to zero to locate the stationary points, we find∂f∂x =(3x2 − 2x 4 )exp(−x 2 − y 2 )=0, (5.23)∂f∂y = −2yx3 exp(−x 2 − y 2 )=0. (5.24)For (5.24) to be satisfied we require x =0ory = 0 <strong>and</strong> <strong>for</strong> (5.23) to be satisfied we requirex =0orx = ± √ 3/2. Hence the stationary points are at (0, 0), ( √ 3/2, 0) <strong>and</strong> (− √ 3/2, 0).We now find the second partial derivatives:f xx =(4x 5 − 14x 3 +6x)exp(−x 2 − y 2 ),f yy = x 3 (4y 2 − 2) exp(−x 2 − y 2 ),f xy =2x 2 y(2x 2 − 3) exp(−x 2 − y 2 ).We then substitute the pairs of values of x <strong>and</strong> y <strong>for</strong> each stationary point <strong>and</strong> find thatat (0, 0)f xx =0, f yy =0, f xy =0<strong>and</strong> at (± √ 3/2, 0)f xx = ∓6 √ 3/2 exp(−3/2), f yy = ∓3 √ 3/2 exp(−3/2), f xy =0.Hence, applying criteria (i)–(iii) above, we find that (0, 0) is an undetermined stationarypoint, ( √ 3/2, 0) is a maximum <strong>and</strong> (− √ 3/2, 0) is a minimum. The function is shown infigure 5.3. ◭Determining the nature of stationary points <strong>for</strong> functions of a general numberof variables is considerably more difficult <strong>and</strong> requires a knowledge of theeigenvectors <strong>and</strong> eigenvalues of matrices. Although these are not discussed untilchapter 8, we present the analysis here <strong>for</strong> completeness. The remainder of thissection can there<strong>for</strong>e be omitted on a first reading.For a function of n real variables, f(x 1 ,x 2 ,...,x n ), we require that, at allstationary points,∂f= 0 <strong>for</strong> all x i .∂x iIn order to determine the nature of a stationary point, we must exp<strong>and</strong> thefunction as a Taylor series about the point. Recalling the Taylor expansion (5.20)<strong>for</strong> a function of n variables, we see that∆f = f(x) − f(x 0 ) ≈ 1 ∑ ∑ ∂ 2 f∆x i ∆x j . (5.25)2 ∂x i ∂x j165ij


PARTIAL DIFFERENTIATIONmaximum0.40.20−0.2−0.4−3−2minimum−1 0x1 2 320y−2Figure 5.3 The function f(x, y) =x 3 exp(−x 2 − y 2 ).If we define the matrix M to have elements given bythen we can rewrite (5.25) asM ij =∂2 f∂x i ∂x j,∆f = 1 2 ∆xT M∆x, (5.26)where ∆x is the column vector with the ∆x i as its components <strong>and</strong> ∆x T is itstranspose. Since M is real <strong>and</strong> symmetric it has n real eigenvalues λ r <strong>and</strong> northogonal eigenvectors e r , which after suitable normalisation satisfyMe r = λ r e r , e T r e s = δ rs ,where the Kronecker delta, written δ rs , equals unity <strong>for</strong> r = s <strong>and</strong> equals zerootherwise. These eigenvectors <strong>for</strong>m a basis set <strong>for</strong> the n-dimensional space <strong>and</strong>we can there<strong>for</strong>e exp<strong>and</strong> ∆x in terms of them, obtaining∆x = ∑ ra r e r ,166


5.9 STATIONARY VALUES UNDER CONSTRAINTSwhere the a r are coefficients dependent upon ∆x. Substituting this into (5.26), wefind∑∆f = 1 2 ∆xT M∆x = 1 2λ r a 2 r .Now, <strong>for</strong> the stationary point to be a minimum, we require ∆f = 1 ∑2 r λ ra 2 r > 0<strong>for</strong> all sets of values of the a r , <strong>and</strong> there<strong>for</strong>e all the eigenvalues of M to begreater than zero. Conversely, <strong>for</strong> a maximum we require ∆f = 1 ∑2 r λ ra 2 r < 0,<strong>and</strong> there<strong>for</strong>e all the eigenvalues of M to be less than zero. If the eigenvalues havemixed signs, then we have a saddle point. Note that the test may fail if some orall of the eigenvalues are equal to zero <strong>and</strong> all the non-zero ones have the samesign.◮Derive the conditions <strong>for</strong> maxima, minima <strong>and</strong> saddle points <strong>for</strong> a function of two realvariables, using the above analysis.For a two-variable function the matrix M is given by( )fxx fM =xy.f yx f yyThere<strong>for</strong>e its eigenvalues satisfy the equation∣ f xx − λ f xyf xy f yy − λ ∣ =0.Hence(f xx − λ)(f yy − λ) − fxy 2 =0⇒ f xx f yy − (f xx + f yy )λ + λ 2 − fxy 2 =0√⇒ 2λ =(f xx + f yy ) ± (f xx + f yy ) 2 − 4(f xx f yy − fxy),2which by rearrangement of the terms under the square root gives√2λ =(f xx + f yy ) ± (f xx − f yy ) 2 +4fxy.2Now, that M is real <strong>and</strong> symmetric implies that its eigenvalues are real, <strong>and</strong> so <strong>for</strong> botheigenvalues to be positive (corresponding to a minimum), we require f xx <strong>and</strong> f yy positive<strong>and</strong> also√f xx + f yy > (f xx + f yy ) 2 − 4(f xx f yy − fxy),2⇒ f xx f yy − fxy 2 > 0.A similar procedure will find the criteria <strong>for</strong> maxima <strong>and</strong> saddle points. ◭r5.9 Stationary values under constraintsIn the previous section we looked at the problem of finding stationary values ofa function of two or more variables when all the variables may be independently167


PARTIAL DIFFERENTIATIONvaried. However, it is often the case in physical problems that not all the variablesused to describe a situation are in fact independent, i.e. some relationshipbetween the variables must be satisfied. For example, if we walk through a hillyl<strong>and</strong>scape <strong>and</strong> we are constrained to walk along a path, we will never reachthe highest peak on the l<strong>and</strong>scape unless the path happens to take us to it.Nevertheless, we can still find the highest point that we have reached during ourjourney.We first discuss the case of a function of just two variables. Let us considerfinding the maximum value of the differentiable function f(x, y) subject to theconstraint g(x, y) =c, wherec is a constant. In the above analogy, f(x, y) mightrepresent the height of the l<strong>and</strong> above sea-level in some hilly region, whilstg(x, y) =c is the equation of the path along which we walk.We could, of course, use the constraint g(x, y) =c to substitute <strong>for</strong> x or y inf(x, y), thereby obtaining a new function of only one variable whose stationarypoints could be found using the methods discussed in subsection 2.1.8. However,such a procedure can involve a lot of algebra <strong>and</strong> becomes very tedious <strong>for</strong> functionsof more than two variables. A more direct method <strong>for</strong> solving such problemsis the method of Lagrange undetermined multipliers, which we now discuss.To maximise f we requiredf = ∂f ∂fdx + dy =0.∂x ∂yIf dx <strong>and</strong> dy were independent, we could conclude f x =0=f y . However, herethey are not independent, but constrained because g is constant:dg = ∂g ∂gdx + dy =0.∂x ∂yMultiplying dg by an as yet unknown number λ <strong>and</strong> adding it to df we obtaind(f + λg) =( ∂f∂x + λ ∂g∂x)dx +( ∂f∂y + λ∂g ∂y)dy =0,where λ is called a Lagrange undetermined multiplier. In this equation dx <strong>and</strong> dyare to be independent <strong>and</strong> arbitrary; we must there<strong>for</strong>e choose λ such that∂f∂x + λ ∂g =0, (5.27)∂x∂f∂y + λ∂g =0. (5.28)∂yThese equations, together with the constraint g(x, y) =c, are sufficient to find thethree unknowns, i.e. λ <strong>and</strong> the values of x <strong>and</strong> y at the stationary point.168


5.9 STATIONARY VALUES UNDER CONSTRAINTS◮The temperature of a point (x, y) on a unit circle is given by T (x, y) =1+xy. Findthetemperature of the two hottest points on the circle.We need to maximise T (x, y) subject to the constraint x 2 + y 2 = 1. Applying (5.27) <strong>and</strong>(5.28), we obtainy +2λx =0, (5.29)x +2λy =0. (5.30)These results, together with the original constraint x 2 + y 2 = 1, provide three simultaneousequations that may be solved <strong>for</strong> λ, x <strong>and</strong> y.From (5.29) <strong>and</strong> (5.30) we find λ = ±1/2, which in turn implies that y = ∓x. Rememberingthat x 2 + y 2 = 1, we find thaty = x ⇒ x = ± √ 1 ,2y = ± √ 12y = −x ⇒ x = ∓ √ 1 ,2y = ± √ 1 .2We have not yet determined which of these stationary points are maxima <strong>and</strong> which areminima. In this simple case, we need only substitute the four pairs of x- <strong>and</strong>y- values intoT (x, y) =1+xy to find that the maximum temperature on the unit circle is T max =3/2 atthe points y = x = ±1/ √ 2. ◭The method of Lagrange multipliers can be used to find the stationary points offunctions of more than two variables, subject to several constraints, provided thatthe number of constraints is smaller than the number of variables. For example,if we wish to find the stationary points of f(x, y, z) subject to the constraintsg(x, y, z) =c 1 <strong>and</strong> h(x, y, z) =c 2 ,wherec 1 <strong>and</strong> c 2 are constants, then we proceedas above, obtaining∂(f + λg + µh) =∂f∂x ∂x + λ ∂g∂x + µ ∂h∂x = 0,∂(f + λg + µh) =∂f∂y ∂y + λ∂g ∂y + µ ∂h∂y= 0, (5.31)∂(f + λg + µh) =∂f∂z ∂z + λ∂g ∂z + µ∂h ∂z = 0.We may now solve these three equations, together with the two constraints, togive λ, µ, x, y <strong>and</strong> z.169


PARTIAL DIFFERENTIATION◮Find the stationary points of f(x, y, z) =x 3 + y 3 + z 3 subject to the following constraints:(i) g(x, y, z) =x 2 + y 2 + z 2 =1;(ii) g(x, y, z) =x 2 + y 2 + z 2 =1 <strong>and</strong> h(x, y, z) =x + y + z =0.Case (i). Since there is only one constraint in this case, we need only introduce a singleLagrange multiplier to obtain∂∂x (f + λg) =3x2 +2λx =0,∂∂y (f + λg) =3y2 +2λy =0, (5.32)∂∂z (f + λg) =3z2 +2λz =0.These equations are highly symmetrical <strong>and</strong> clearly have the solution x = y = z = −2λ/3.Using the constraint x 2 + y 2 + z 2 = 1 we find λ = ± √ 3/2 <strong>and</strong> so stationary points occuratx = y = z = ± √ 1 . (5.33)3In solving the three equations (5.32) in this way, however, we have implicitly assumedthat x, y <strong>and</strong> z are non-zero. However, it is clear from (5.32) that any of these values canequal zero, with the exception of the case x = y = z = 0 since this is prohibited by theconstraint x 2 + y 2 + z 2 = 1. We must consider the other cases separately.If x = 0, <strong>for</strong> example, we require3y 2 +2λy =0,3z 2 +2λz =0,y 2 + z 2 =1.Clearly, we require λ ≠ 0, otherwise these equations are inconsistent. If neither y norz is zero we find y = −2λ/3 =z <strong>and</strong> from the third equation we require y = z =±1/ √ 2. If y =0,however,thenz = ±1 <strong>and</strong>, similarly, if z =0theny = ±1. Thus thestationary points having x =0are(0, 0, ±1), (0, ±1, 0) <strong>and</strong> (0, ±1/ √ 2, ±1/ √ 2). A similarprocedure can be followed <strong>for</strong> the cases y =0<strong>and</strong>z = 0 respectively <strong>and</strong>, in additionto those already obtained, we find the stationary points (±1, 0, 0), (±1/ √ 2, 0, ±1/ √ 2) <strong>and</strong>(±1/ √ 2, ±1/ √ 2, 0).Case (ii). We now have two constraints <strong>and</strong> must there<strong>for</strong>e introduce two Lagrangemultipliers to obtain (cf. (5.31))∂∂x (f + λg + µh) =3x2 +2λx + µ =0, (5.34)∂∂y (f + λg + µh) =3y2 +2λy + µ =0, (5.35)∂∂z (f + λg + µh) =3z2 +2λz + µ =0. (5.36)These equations are again highly symmetrical <strong>and</strong> the simplest way to proceed is tosubtract (5.35) from (5.34) to obtain3(x 2 − y 2 )+2λ(x − y) =0⇒ 3(x + y)(x − y)+2λ(x − y) =0. (5.37)This equation is clearly satisfied if x = y; then, from the second constraint, x + y + z =0,170


5.9 STATIONARY VALUES UNDER CONSTRAINTSwe find z = −2x. Substituting these values into the first constraint, x 2 + y 2 + z 2 =1,weobtainx = ± √ 1 , y = ± √ 1 , z = ∓ √ 2 . (5.38)6 6 6Because of the high degree of symmetry amongst the equations (5.34)–(5.36), we may obtainby inspection two further relations analogous to (5.37), one containing the variables y, z<strong>and</strong> the other the variables x, z. Assuming y = z in the first relation <strong>and</strong> x = z in thesecond, we find the stationary pointsx = ± 1 √6, y = ∓ 2 √6, z = ± 1 √6(5.39)<strong>and</strong>x = ∓ 2 √6, y = ± 1 √6, z = ± 1 √6. (5.40)We note that in finding the stationary points (5.38)–(5.40) we did not need to evaluate theLagrange multipliers λ <strong>and</strong> µ explicitly. This is not always the case, however, <strong>and</strong> in someproblems it may be simpler to begin by finding the values of these multipliers.Returning to (5.37) we must now consider the case where x ≠ y; then we find3(x + y)+2λ =0. (5.41)However, in obtaining the stationary points (5.39), (5.40), we did not assume x = y butonly required y = z <strong>and</strong> x = z respectively. It is clear that x ≠ y at these stationary points,<strong>and</strong> it can be shown that they do indeed satisfy (5.41). Similarly, several stationary points<strong>for</strong> which x ≠ z or y ≠ z have already been found.Thus we need to consider further only two cases, x = y = z, <strong>and</strong>x, y <strong>and</strong> z are alldifferent. The first is clearly prohibited by the constraint x + y + z = 0. For the secondcase, (5.41) must be satisfied, together with the analogous equations containing y, z <strong>and</strong>x, z respectively, i.e.3(x + y)+2λ =0,3(y + z)+2λ =0,3(x + z)+2λ =0.Adding these three equations together <strong>and</strong> using the constraint x+y +z = 0 we find λ =0.However, <strong>for</strong> λ = 0 the equations are inconsistent <strong>for</strong> non-zero x, y <strong>and</strong> z. There<strong>for</strong>e allthe stationary points have already been found <strong>and</strong> are given by (5.38)–(5.40). ◭The method may be extended to functions of any number n of variablessubject to any smaller number m of constraints. This means that effectively thereare n − m independent variables <strong>and</strong>, as mentioned above, we could solve bysubstitution <strong>and</strong> then by the methods of the previous section. However, <strong>for</strong> largen this becomes cumbersome <strong>and</strong> the use of Lagrange undetermined multipliers isa useful simplification.171


PARTIAL DIFFERENTIATION◮A system contains a very large number N of particles, each of which can be in any of Renergy levels with a corresponding energy E i , i =1, 2,...,R. The number of particles in theith level is n i <strong>and</strong> the total energy of the system is a constant, E. Find the distribution ofparticles amongst the energy levels that maximises the expressionN!P =n 1 !n 2 ! ···n R ! ,subject to the constraints that both the number of particles <strong>and</strong> the total energy remainconstant, i.e.R∑R∑g = N − n i =0 <strong>and</strong> h = E − n i E i =0.i=1i=1The way in which we proceed is as follows. In order to maximise P , we must minimiseits denominator (since the numerator is fixed). Minimising the denominator is the same asminimising the logarithm of the denominator, i.e.f =ln(n 1 !n 2 ! ···n R !) =ln(n 1 !) +ln(n 2 !) + ···+ln(n R !) .Using Stirling’s approximation, ln (n!) ≈ n ln n − n, we find thatf = n 1 ln n 1 + n 2 ln n 2 + ···+ n R ln n R − (n 1 + n 2 + ···+ n R )( R∑)= n i ln n i − N.i=1It has been assumed here that, <strong>for</strong> the desired distribution, all the n i are large. Thus, wenow have a function f subject to two constraints, g =0<strong>and</strong>h = 0, <strong>and</strong> we can apply theLagrange method, obtaining (cf. (5.31))∂f+ λ ∂g + µ ∂h =0,∂n 1 ∂n 1 ∂n 1∂f+ λ ∂g + µ ∂h =0,∂n 2 ∂n 2 ∂n 2 .∂f+ λ ∂g + µ ∂h =0.∂n R ∂n R ∂n RSince all these equations are alike, we consider the general case∂f+ λ ∂g + µ ∂h =0,∂n k ∂n k ∂n k<strong>for</strong> k =1, 2,...,R. Substituting the functions f, g <strong>and</strong> h into this relation we findn k+lnn k + λ(−1) + µ(−E k )=0,n kwhich can be rearranged to giveln n k = µE k + λ − 1,<strong>and</strong> hencen k = C exp µE k .172


5.10 ENVELOPESWe now have the general <strong>for</strong>m <strong>for</strong> the distribution of particles amongst energy levels, butin order to determine the two constants µ, C we recall that<strong>and</strong>R∑C exp µE k = Nk=1R∑CE k exp µE k = E.k=1This is known as the Boltzmann distribution <strong>and</strong> is a well-known result from statisticalmechanics. ◭5.10 EnvelopesAs noted at the start of this chapter, many of the functions with which physicists,chemists <strong>and</strong> engineers have to deal contain, in addition to constants <strong>and</strong> oneor more variables, quantities that are normally considered as parameters of thesystem under study. Such parameters may, <strong>for</strong> example, represent the capacitanceof a capacitor, the length of a rod, or the mass of a particle – quantities thatare normally taken as fixed <strong>for</strong> any particular physical set-up. The correspondingvariables may well be time, currents, charges, positions <strong>and</strong> velocities. However,the parameters could be varied <strong>and</strong> in this section we study the effects of doing so;in particular we study how the <strong>for</strong>m of dependence of one variable on another,typically y = y(x), is affected when the value of a parameter is changed in asmooth <strong>and</strong> continuous way. In effect, we are making the parameter into anadditional variable.As a particular parameter, which we denote by α, is varied over its permittedrange, the shape of the plot of y against x will change, usually, but not always,in a smooth <strong>and</strong> continuous way. For example, if the muzzle speed v of a shellfired from a gun is increased through a range of values then its height–distancetrajectories will be a series of curves with a common starting point that areessentially just magnified copies of the original; furthermore the curves do notcross each other. However, if the muzzle speed is kept constant but θ, the angleof elevation of the gun, is increased through a series of values, the correspondingtrajectories do not vary in a monotonic way. When θ has been increased beyond45 ◦ the trajectories then do cross some of the trajectories corresponding to θ45 ◦ all lie within a curve that touches each individualtrajectory at one point. Such a curve is called the envelope to the set of trajectorysolutions; it is to the study of such envelopes that this section is devoted.For our general discussion of envelopes we will consider an equation of the<strong>for</strong>m f = f(x, y, α) = 0. A function of three Cartesian variables, f = f(x, y, α),is defined at all points in xyα-space, whereas f = f(x, y, α) =0isasurface inthis space. A plane of constant α, which is parallel to the xy-plane, cuts such173


PARTIAL DIFFERENTIATIONP 1PP 2yf(x, y, α 1 )=0 f(x, y, α 1 + h) =0xFigure 5.4 Two neighbouring curves in the xy-plane of the family f(x, y, α) =0 intersecting at P .Forfixedα 1 , the point P 1 is the limiting position of P ash → 0. As α 1 is varied, P 1 delineates the envelope of the family (broken line).a surface in a curve. Thus different values of the parameter α correspond todifferent curves, which can be plotted in the xy-plane. We now investigate howthe envelope equation <strong>for</strong> such a family of curves is obtained.5.10.1 Envelope equationsSuppose f(x, y, α 1 )=0<strong>and</strong>f(x, y, α 1 + h) = 0 are two neighbouring curves of afamily <strong>for</strong> which the parameter α differs by a small amount h. Let them intersectat the point P with coordinates x, y, as shown in figure 5.4. Then the envelope,indicated by the broken line in the figure, touches f(x, y, α 1 ) = 0 at the point P 1 ,which is defined as the limiting position of P when α 1 is fixed but h → 0. Thefull envelope is the curve traced out by P 1 as α 1 changes to generate successivemembers of the family of curves. Of course, <strong>for</strong> any finite h, f(x, y, α 1 + h) =0isone of these curves <strong>and</strong> the envelope touches it at the point P 2 .We are now going to apply Rolle’s theorem, see subsection 2.1.10, with theparameter α as the independent variable <strong>and</strong> x <strong>and</strong> y fixed as constants. In thiscontext, the two curves in figure 5.4 can be thought of as the projections onto thexy-plane of the planar curves in which the surface f = f(x, y, α) = 0 meets theplanes α = α 1 <strong>and</strong> α = α 1 + h.Along the normal to the page that passes through P ,asα changes from α 1to α 1 + h the value of f = f(x, y, α) will depart from zero, because the normalmeets the surface f = f(x, y, α) =0onlyatα = α 1 <strong>and</strong> at α = α 1 + h. However,at these end points the values of f = f(x, y, α) will both be zero, <strong>and</strong> there<strong>for</strong>eequal. This allows us to apply Rolle’s theorem <strong>and</strong> so to conclude that <strong>for</strong> someθ in the range 0 ≤ θ ≤ 1 the partial derivative ∂f(x, y, α 1 + θh)/∂α is zero. When174


5.10 ENVELOPESh is made arbitrarily small, so that P → P 1 , the three defining equations reduceto two, which define the envelope point P 1 :∂f(x, y, α 1 )f(x, y, α 1 )=0 <strong>and</strong>=0. (5.42)∂αIn (5.42) both the function <strong>and</strong> the gradient are evaluated at α = α 1 . The equationof the envelope g(x, y) = 0 is found by eliminating α 1 between the two equations.As a simple example we will now solve the problem which when posed mathematicallyreads ‘calculate the envelope appropriate to the family of straight linesin the xy-plane whose points of intersection with the coordinate axes are a fixeddistance apart’. In more ordinary language, the problem is about a ladder leaningagainst a wall.◮A ladder of length L st<strong>and</strong>s on level ground <strong>and</strong> can be leaned at any angle against avertical wall. Find the equation of the curve bounding the vertical area below the ladder.We take the ground <strong>and</strong> the wall as the x- <strong>and</strong>y-axes respectively. If the foot of the ladderis a from the foot of the wall <strong>and</strong> the top is b above the ground then the straight-lineequation of the ladder isxa + y b =1,where a <strong>and</strong> b are connected by a 2 + b 2 = L 2 . Expressed in st<strong>and</strong>ard <strong>for</strong>m with only oneindependent parameter, a, the equation becomesf(x, y, a) = x a + y− 1=0. (5.43)(L 2 − a 2 )1/2Now, differentiating (5.43) with respect to a <strong>and</strong> setting the derivative ∂f/∂a equal tozero gives− x a + ay=0;2 (L 2 − a 2 )3/2from which it follows thatLx 1/3a =<strong>and</strong> (L 2 − a 2 ) 1/2 Ly 1/3=(x 2/3 + y 2/3 ) 1/2 (x 2/3 + y 2/3 ) . 1/2Eliminating a by substituting these values into (5.43) gives, <strong>for</strong> the equation of theenvelope of all possible positions on the ladder,x 2/3 + y 2/3 = L 2/3 .This is the equation of an astroid (mentioned in exercise 2.19), <strong>and</strong>, together with the wall<strong>and</strong> the ground, marks the boundary of the vertical area below the ladder. ◭Other examples, drawn from both geometry <strong>and</strong> <strong>and</strong> the physical sciences, areconsidered in the exercises at the end of this chapter. The shell trajectory problemdiscussed earlier in this section is solved there, but in the guise of a questionabout the water bell of an ornamental fountain.175


PARTIAL DIFFERENTIATION5.11 Thermodynamic relationsThermodynamic relations provide a useful set of physical examples of partialdifferentiation. The relations we will derive are called Maxwell’s thermodynamicrelations. They express relationships between four thermodynamic quantities describinga unit mass of a substance. The quantities are the pressure P , the volumeV , the thermodynamic temperature T <strong>and</strong> the entropy S of the substance. Thesefour quantities are not independent; any two of them can be varied independently,but the other two are then determined.The first law of thermodynamics may be expressed asdU = TdS− PdV, (5.44)where U is the internal energy of the substance. Essentially this is a conservationof energy equation, but we shall concern ourselves, not with the physics, but ratherwith the use of partial differentials to relate the four basic quantities discussedabove. The method involves writing a total differential, dU say, in terms of thedifferentials of two variables, say X <strong>and</strong> Y , thus( ) ( )∂U∂UdU = dX + dY , (5.45)∂XY∂YX<strong>and</strong> then using the relationship∂ 2 U∂X∂Y =∂2 U∂Y ∂Xto obtain the required Maxwell relation. The variables X <strong>and</strong> Y aretobechosenfrom P , V , T <strong>and</strong> S.◮Show that (∂T/∂V) S = −(∂P/∂S) V .Here the two variables that have to be held constant, in turn, happen to be those whosedifferentials appear on the RHS of (5.44). And so, taking X as S <strong>and</strong> Y as V in (5.45), wehave<strong>and</strong> find directly thatTdS− PdV= dU =( ) ∂U= T <strong>and</strong>∂SV( ) ∂UdS +∂SV( ) ∂UdV ,∂VS( ) ∂U= −P.∂VSDifferentiating the first expression with respect to V <strong>and</strong> the second with respect to S, <strong>and</strong>using∂ 2 U∂V∂S =∂2 U∂S∂V ,we find the Maxwell relation( ) ( )∂T ∂P= − . ◭∂VS∂SV176


5.11 THERMODYNAMIC RELATIONS◮Show that (∂S/∂V) T =(∂P/∂T) V .Applying (5.45) to dS, with independent variables V <strong>and</strong> T , we find[( ) ( ) ]∂S∂SdU = TdS− PdV= T dV + dT − PdV.∂VT∂TVSimilarly applying (5.45) to dU, we find( ) ∂UdU =∂VThus, equating partial derivatives,( ) ( )∂U ∂S= T − P∂V ∂VBut, sinceT∂ 2 U∂T∂V =TdV +T<strong>and</strong>∂2 U∂V∂T , i.e. ∂∂Tit follows that( ) ∂S∂VTThus finally we get the Maxwell relation( ) ∂S∂V( ) ∂UdT .∂TV( ) ( )∂U ∂S= T .∂TV∂TV( ) ∂U= ∂∂VT∂V( ) ∂U,∂TV( )+ T ∂2 S ∂P∂T∂V − = ∂ [ ( ) ] ∂ST= T ∂2 S∂TV∂V ∂TV T∂V∂T .T=( ) ∂P. ◭∂TVThe above derivation is rather cumbersome, however, <strong>and</strong> a useful trick thatcan simplify the working is to define a new function, called a potential. Theinternal energy U discussed above is one example of a potential but three othersare commonly defined <strong>and</strong> they are described below.◮Show that (∂S/∂V) T =(∂P/∂T) V by considering the potential U − ST.We first consider the differential d(U − ST). From (5.5), we obtaind(U − ST)=dU − SdT − TdS = −SdT − PdVwhen use is made of (5.44). We rewrite U − ST as F <strong>for</strong> convenience of notation; F iscalled the Helmholtz potential. ThusdF = −SdT − PdV,<strong>and</strong> it follows that( )( )∂F∂F= −S <strong>and</strong>= −P.∂TV∂VTUsing these results together with∂ 2 F∂T∂V =∂2 F∂V∂T ,we can see immediately that( ) ( )∂S ∂P= ,∂VT∂TVwhich is the same Maxwell relation as be<strong>for</strong>e. ◭177


PARTIAL DIFFERENTIATIONAlthough the Helmholtz potential has other uses, in this context it has simplyprovided a means <strong>for</strong> a quick derivation of the Maxwell relation. The otherMaxwell relations can be derived similarly by using two other potentials, theenthalpy, H = U + PV, <strong>and</strong> the Gibbs free energy, G = U + PV − ST (seeexercise 5.25).5.12 Differentiation of integralsWe conclude this chapter with a discussion of the differentiation of integrals. Letus consider the indefinite integral (cf. equation (2.30))∫F(x, t) = f(x, t) dt,from which it follows immediately that∂F(x, t)= f(x, t).∂tAssuming that the second partial derivatives of F(x, t) are continuous, we have∂ 2 F(x, t)∂t∂x= ∂2 F(x, t),∂x∂t<strong>and</strong> so we can write[ ]∂ ∂F(x, t)= ∂ [ ] ∂F(x, t) ∂f(x, t)=∂t ∂x ∂x ∂t ∂x .Integrating this equation with respect to t then gives∫∂F(x, t) ∂f(x, t)= dt. (5.46)∂x ∂xNow consider the definite integralI(x) =∫ t=vt=uf(x, t) dt= F(x, v) − F(x, u),where u <strong>and</strong> v are constants. Differentiating this integral with respect to x, <strong>and</strong>using (5.46), we see thatdI(x) ∂F(x, v)=dx ∂x=u∂F(x, u)∂x−∫ v ∫∂f(x, t) u∂x dt − ∂f(x, t)∂xdt∫ v∂f(x, t)=∂xdt.This is Leibnitz’ rule <strong>for</strong> differentiating integrals, <strong>and</strong> basically it states that <strong>for</strong>178


5.13 EXERCISESconstant limits of integration the order of integration <strong>and</strong> differentiation can bereversed.In the more general case where the limits of the integral are themselves functionsof x, it follows immediately thatI(x) =∫ t=v(x)t=u(x)which yields the partial derivatives∂I∂I= f(x, v(x)),∂vConsequentlyf(x, t) dt= F(x, v(x)) − F(x, u(x)),( ) ( )dI ∂I dv ∂I dudx = ∂v dx + ∂u dx + ∂I∂x= −f(x, u(x)).∂u= f(x, v(x)) dvdu− f(x, u(x))dx dx + ∂∂x= f(x, v(x)) dv∫du v(x)− f(x, u(x))dx dx + u(x)∫ v(x)u(x)f(x, t)dt∂f(x, t)∂xdt, (5.47)where the partial derivative with respect to x in the last term has been takeninside the integral sign using (5.46). This procedure is valid because u(x) <strong>and</strong>v(x)are being held constant in this term.◮Find the derivative with respect to x of the integralApplying (5.47), we see thatdIdxI(x) =∫ x 2sin x3 sin x2= (2x) −x 2= 2sinx3x−xsin x2xsin xtdt.t∫ x 2x (1) + [ sin xt+xsin x3 sin x2=3 − 2x x= 1 x (3 sin x3 − 2sinx 2 ). ◭x] x 2xt cos xtdtt5.13 Exercises5.1 Using the appropriate properties of ordinary derivatives, per<strong>for</strong>m the following.179


PARTIAL DIFFERENTIATION(a) Find all the first partial derivatives of the following functions f(x, y):(i) x 2 y, (ii) x 2 + y 2 + 4, (iii) sin(x/y), (iv) tan −1 (y/x),(v) r(x, y, z) =(x 2 + y 2 + z 2 ) 1/2 .(b) For (i), (ii) <strong>and</strong> (v), find ∂ 2 f/∂x 2 , ∂ 2 f/∂y 2 <strong>and</strong> ∂ 2 f/∂x∂y.(c) For (iv) verify that ∂ 2 f/∂x∂y = ∂ 2 f/∂y∂x.5.2 Determine which of the following are exact differentials:(a) (3x +2)ydx+ x(x +1)dy;(b) y tan xdx+ x tan ydy;(c) y 2 (ln x +1)dx +2xy ln xdy;(d) y 2 (ln x +1)dy +2xy ln xdx;(e) [x/(x 2 + y 2 )] dy − [y/(x 2 + y 2 )] dx.5.3 Show that the differentialdf = x 2 dy − (y 2 + xy) dxis not exact, but that dg =(xy 2 ) −1 df is exact.5.4 Show thatdf = y(1 + x − x 2 ) dx + x(x +1)dyis not an exact differential.Find the differential equation that a function g(x) mustsatisfyifdφ = g(x)dfis to be an exact differential. Verify that g(x) =e −x is a solution of this equation<strong>and</strong> deduce the <strong>for</strong>m of φ(x, y).5.5 The equation 3y = z 3 +3xz defines z implicitly as a function of x <strong>and</strong> y. Evaluateall three second partial derivatives of z with respect to x <strong>and</strong>/or y. Verifythatzis a solution ofx ∂2 z∂y + ∂2 z2 ∂x =0. 25.6 A possible equation of state <strong>for</strong> a gas takes the <strong>for</strong>m(PV = RT exp −α ),VRTin which α <strong>and</strong> R are constants. Calculate expressions <strong>for</strong>( ) ( ) ( )∂P∂V∂T∂V∂T∂P,T<strong>and</strong> show that their product is −1, as stated in section 5.4.5.7 The function G(t) is defined byG(t) =F(x, y) =x 2 + y 2 +3xy,where x(t) =at 2 <strong>and</strong> y(t) =2at. Use the chain rule to find the values of (x, y) atwhich G(t) has stationary values as a function of t. Do any of them correspondto the stationary points of F(x, y) as a function of x <strong>and</strong> y?5.8 In the xy-plane, new coordinates s <strong>and</strong> t are defined bys = 1 (x + y), t 2Trans<strong>for</strong>m the equation,P= 1 (x − y).2∂ 2 φ∂x − ∂2 φ2 ∂y =0 2into the new coordinates <strong>and</strong> deduce that its general solution can be writtenφ(x, y) =f(x + y)+g(x − y),where f(u) <strong>and</strong>g(v) are arbitrary functions of u <strong>and</strong> v, respectively.180,V


5.13 EXERCISES5.9 The function f(x, y) satisfies the differential equationy ∂f∂x + x ∂f∂y =0.By changing to new variables u = x 2 − y 2 <strong>and</strong> v =2xy, show that f is, in fact, afunction of x 2 − y 2 only.5.10 If x = e u cos θ <strong>and</strong> y = e u sin θ, show that( )∂ 2 φ∂u + ∂2 φ∂ 2 ∂θ 2 =(x2 + y 2 2 f)∂x + ∂2 f,2 ∂y 2where f(x, y) =φ(u, θ).5.11 Find <strong>and</strong> evaluate the maxima, minima <strong>and</strong> saddle points of the functionf(x, y) =xy(x 2 + y 2 − 1).5.12 Show thatf(x, y) =x 3 − 12xy +48x + by 2 , b ≠0,has two, one, or zero stationary points, according to whether |b| is less than,equal to, or greater than 3.5.13 Locate the stationary points of the functionf(x, y) =(x 2 − 2y 2 )exp[−(x 2 + y 2 )/a 2 ],where a is a non-zero constant.Sketch the function along the x- <strong>and</strong>y-axes <strong>and</strong> hence identify the nature <strong>and</strong>values of the stationary points.5.14 Find the stationary points of the functionf(x, y) =x 3 + xy 2 − 12x − y 2<strong>and</strong> identify their natures.5.15 Find the stationary values off(x, y) =4x 2 +4y 2 + x 4 − 6x 2 y 2 + y 4<strong>and</strong> classify them as maxima, minima or saddle points. Make a rough sketch ofthe contours of f in the quarter plane x, y ≥ 0.5.16 The temperature of a point (x, y, z) on the unit sphere is given byT (x, y, z) =1+xy + yz.By using the method of Lagrange multipliers, find the temperature of the hottestpoint on the sphere.5.17 A rectangular parallelepiped has all eight vertices on the ellipsoidx 2 +3y 2 +3z 2 =1.Using the symmetry of the parallelepiped about each of the planes x = 0,y =0,z = 0, write down the surface area of the parallelepiped in terms ofthe coordinates of the vertex that lies in the octant x, y, z ≥ 0. Hence find themaximum value of the surface area of such a parallelepiped.5.18 Two horizontal corridors, 0 ≤ x ≤ a with y ≥ 0, <strong>and</strong> 0 ≤ y ≤ b with x ≥ 0, meetat right angles. Find the length L of the longest ladder (considered as a stick)that may be carried horizontally around the corner.5.19 A barn is to be constructed with a uni<strong>for</strong>m cross-sectional area A throughoutits length. The cross-section is to be a rectangle of wall height h (fixed) <strong>and</strong>width w, surmounted by an isosceles triangular roof that makes an angle θ with181


PARTIAL DIFFERENTIATIONthe horizontal. The cost of construction is α perunitheightofwall<strong>and</strong>β perunit (slope) length of roof. Show that, irrespective of the values of α <strong>and</strong> β, tominimise costs w should be chosen to satisfy the equationw 4 =16A(A − wh),<strong>and</strong> θ made such that 2 tan 2θ = w/h.5.20 Show that the envelope of all concentric ellipses that have their axes along thex- <strong>and</strong>y-coordinate axes, <strong>and</strong> that have the sum of their semi-axes equal to aconstant L, is the same curve (an astroid) as that found in the worked examplein section 5.10.5.21 Find the area of the region covered by points on the linesxa + y b =1,where the sum of any line’s intercepts on the coordinate axes is fixed <strong>and</strong> equalto c.5.22 Prove that the envelope of the circles whose diameters are those chords of agiven circle that pass through a fixed point on its circumference, is the cardioidr = a(1 + cos θ).Here a is the radius of the given circle <strong>and</strong> (r, θ) are the polar coordinates of theenvelope. Take as the system parameter the angle φ between a chord <strong>and</strong> thepolar axis from which θ is measured.5.23 A water feature contains a spray head at water level at the centre of a roundbasin. The head is in the <strong>for</strong>m of a small hemisphere per<strong>for</strong>ated by many evenlydistributed small holes, through which water spurts out at the same speed, v 0 ,inall directions.(a) What is the shape of the ‘water bell’ so <strong>for</strong>med?(b) What must be the minimum diameter of the bowl if no water is to be lost?5.24 In order to make a focussing mirror that concentrates parallel axial rays to onespot (or conversely <strong>for</strong>ms a parallel beam from a point source), a parabolic shapeshould be adopted. If a mirror that is part of a circular cylinder or sphere wereused, the light would be spread out along a curve. This curve is known as acaustic <strong>and</strong> is the envelope of the rays reflected from the mirror. Denoting by θthe angle which a typical incident axial ray makes with the normal to the mirrorat the place where it is reflected, the geometry of reflection (the angle of incidenceequals the angle of reflection) is shown in figure 5.5.Show that a parametric specification of the caustic isx = R cos θ ( 12 +sin2 θ ) , y = R sin 3 θ,where R is the radius of curvature of the mirror. The curve is, in fact, part of anepicycloid.5.25 By considering the differentialdG = d(U + PV − ST),where G is the Gibbs free energy, P the pressure, V the volume, S the entropy<strong>and</strong> T the temperature of a system, <strong>and</strong> given further that the internal energy UsatisfiesdU = TdS− PdV,derive a Maxwell relation connecting (∂V/∂T) P <strong>and</strong> (∂S/∂P) T .182


5.13 EXERCISESyORθθ2θxFigure 5.5 The reflecting mirror discussed in exercise 5.24.5.26 Functions P (V,T), U(V,T)<strong>and</strong>S(V,T) are related byTdS= dU + PdV,where the symbols have the same meaning as in the previous question. Thepressure P is known from experiment to have the <strong>for</strong>min appropriate units. IfP = T 43 + T V ,U = αV T 4 + βT,where α, β, are constants (or, at least, do not depend on T or V ), deduce that αmust have a specific value, but that β may have any value. Find the corresponding<strong>for</strong>m of S.5.27 As in the previous two exercises on the thermodynamics of a simple gas, thequantity dS = T −1 (dU + PdV) is an exact differential. Use this to prove that( ∂U∂V)T= T( ∂P∂T)V− P.In the van der Waals model of a gas, P obeys the equationP =RTV − b − aV , 2where R, a <strong>and</strong> b are constants. Further, in the limit V →∞, the <strong>for</strong>m of Ubecomes U = cT , where c is another constant. Find the complete expression <strong>for</strong>U(V,T).5.28 The entropy S(H,T), the magnetisation M(H,T) <strong>and</strong> the internal energy U(H,T)of a magnetic salt placed in a magnetic field of strength H, at temperature T ,are connected by the equationTdS = dU − HdM.183


PARTIAL DIFFERENTIATIONBy considering d(U − TS − HM) prove that( ) ( ∂M ∂S=∂TH∂HFor a particular salt,M(H,T)=M 0 [1 − exp(−αH/T )].Show that if, at a fixed temperature, the applied field is increased from zero to astrength such that the magnetization of the salt is 3 M 4 0, then the salt’s entropydecreases by an amountM 0(3 − ln 4).4α5.29 Using the results of section 5.12, evaluate the integralHence show that5.30 The integralI(y) =∫ ∞0).Te −xy sin xdx.x∫ ∞sin xJ =0 x dx = π 2 .∫ ∞−∞e −αx2 dxhas the value (π/α) 1/2 . Use this result to evaluateJ(n) =∫ ∞−∞x 2n e −x2 dx,where n is a positive integer. Express your answer in terms of factorials.5.31 The function f(x) is differentiable <strong>and</strong> f(0) = 0. A second function g(y) is definedby∫ yf(x) dxg(y) = √ .0 y − xProve that∫dg ydy = df dx√ .0 dx y − xFor the case f(x) =x n , prove thatd n gdy =2(n!)√ y.n5.32 The functions f(x, t) <strong>and</strong>F(x) are defined byf(x, t) =e −xt ,F(x) =Verify, by explicit calculation, that∫ x∫dFxdx = f(x, x)+1840f(x, t) dt.0∂f(x, t)dt.∂x


5.14 HINTS AND ANSWERS5.33 If∫ 1I(α) =0x α − 1dx, α > −1,ln xwhat is the value of I(0)? Show thatddα xα = x α ln x,<strong>and</strong> deduce thatddα I(α) = 1α +1 .Hence prove that I(α) =ln(1+α).5.34 Find the derivative, with respect to x, of the integralI(x) =∫ 3xxexp xt dt.5.35 The function G(t, ξ) is defined <strong>for</strong> 0 ≤ t ≤ π by{− cos t sin ξ <strong>for</strong> ξ ≤ t,G(t, ξ) =− sin t cos ξ <strong>for</strong> ξ>t.Show that the function x(t) defined byx(t) =∫ π0G(t, ξ)f(ξ) dξsatisfies the equationd 2 xdt + x = f(t),2where f(t) canbeany arbitrary (continuous) function. Show further that x(0) =[dx/dt] t=π = 0, again <strong>for</strong> any f(t), but that the value of x(π) does depend uponthe <strong>for</strong>m of f(t).[ The function G(t, ξ) is an example of a Green’s function, an importantconcept in the solution of differential equations <strong>and</strong> one studied extensively inlater chapters. ]5.14 Hints <strong>and</strong> answers5.1 (a) (i) 2xy, x 2 ; (ii) 2x, 2y; (iii) y −1 cos(x/y), (−x/y 2 )cos(x/y);(iv) −y/(x 2 + y 2 ),x/(x 2 + y 2 ); (v) x/r, y/r, z/r.(b) (i) 2y, 0, 2x; (ii) 2, 2, 0; (v) (y 2 + z 2 )r −3 , (x 2 + z 2 )r −3 , −xyr −3 .(c) Both second derivatives are equal to (y 2 − x 2 )(x 2 + y 2 ) −2 .5.3 2x ≠ −2y − x. For g, both sides of equation (5.9) equal y −2 .5.5 ∂ 2 z/∂x 2 =2xz(z 2 + x) −3 , ∂ 2 z/∂x∂y =(z 2 − x)(z 2 + x) −3 , ∂ 2 z/∂y 2 = −2z(z 2 + x) −3 .5.7 (0, 0), (a/4, −a) <strong>and</strong>(16a, −8a). Only the saddle point at (0, 0).5.9 The trans<strong>for</strong>med equation is 2(x 2 + y 2 )∂f/∂v = 0; hence f does not depend on v.5.11 Maxima, equal to 1/8, at ±(1/2, −1/2), minima, equal to −1/8, at ±(1/2, 1/2),saddle points, equalling 0, at (0, 0), (0, ±1), (±1, 0).5.13 Maxima equal to a 2 e −1 at (±a, 0), minima equal to −2a 2 e −1 at (0, ±a), saddlepoint equalling 0 at (0, 0).5.15 Minimum at (0, 0); saddle points at (±1, ±1). To help with sketching the contours,determine the behaviour of g(x) =f(x, x).5.17 The Lagrange multiplier method gives z = y = x/2, <strong>for</strong> a maximal area of 4.185


PARTIAL DIFFERENTIATION5.19 The cost always includes 2αh, which can there<strong>for</strong>e be ignored in the optimisation.With Lagrange multiplier λ, sinθ = λw/(4β) <strong>and</strong>β sec θ − 1 λw tan θ = λh, leading2to the stated results.5.21 The envelope of the lines x/a + y/(c−a) − 1 = 0, as a is varied, is √ x+ √ y = √ c.Area = c 2 /6.5.23 (a) Using α =cotθ, whereθ is the initial angle a jet makes with the vertical, theequation is f(z, ρ,α) =z−ρα+[gρ 2 (1+α 2 )/(2v0 2 )], <strong>and</strong> setting ∂f/∂α = 0 givesα = v0 2/(gρ). The water bell has a parabolic profile z = v2 0 /(2g) − gρ2 /(2v0 2).(b) Setting z = 0 gives the minimum diameter as 2v0 2/g.5.25 Show that (∂G/∂P) T = V <strong>and</strong> (∂G/∂T) P = −S. From each result, obtain anexpression <strong>for</strong> ∂ 2 G/∂T∂P <strong>and</strong> equate these, giving (∂V/∂T) P = −(∂S/∂P) T .5.27 Find expressions <strong>for</strong> (∂S/∂V) T <strong>and</strong> (∂S/∂T) V , <strong>and</strong> equate ∂ 2 S/∂V∂T with5.29∂ 2 S/∂T∂V. U(V,T)=cT − aV −1 .dI/dy = −Im[ ∫ ∞0 y2 ). Integrate dI/dy from 0 to ∞.I(∞) =0<strong>and</strong>I(0) = J.5.31 Integrate the RHS of the equation by parts, be<strong>for</strong>e differentiating with respectto y. Repeated application of the method establishes the result <strong>for</strong> all orders ofderivative.5.33 I(0) = 0; use Leibnitz’ rule.5.35 Write x(t) =− cos t ∫ tπcos ξf(ξ) dξ <strong>and</strong> differentiate each0 tterm as a product to obtain dx/dt. Obtaind 2 x/dt 2 in a similar way. Note thatintegrals that have equal lower <strong>and</strong> upper limits have value zero. The value ofx(π) is ∫ π0186


6Multiple integralsFor functions of several variables, just as we may consider derivatives with respectto two or more of them, so may the integral of the function with respect to morethan one variable be <strong>for</strong>med. The <strong>for</strong>mal definitions of such multiple integrals areextensions of that <strong>for</strong> a single variable, discussed in chapter 2. We first discussdouble <strong>and</strong> triple integrals <strong>and</strong> illustrate some of their applications. We thenconsider changing the variables in multiple integrals <strong>and</strong> discuss some generalproperties of Jacobians.6.1 Double integralsFor an integral involving two variables – a double integral – we have a function,f(x, y) say, to be integrated with respect to x <strong>and</strong> y between certain limits. Theselimits can usually be represented by a closed curve C bounding a region R in thexy-plane. Following the discussion of single integrals given in chapter 2, let usdivide the region R into N subregions ∆R p of area ∆A p , p =1, 2,...,N, <strong>and</strong> let(x p ,y p ) be any point in subregion ∆R p . Now consider the sumS =N∑f(x p ,y p )∆A p ,p=1<strong>and</strong> let N →∞as each of the areas ∆A p → 0. If the sum S tends to a uniquelimit, I, then this is called the double integral of f(x, y) over the region R <strong>and</strong> iswritten∫I = f(x, y) dA, (6.1)Rwhere dA st<strong>and</strong>s <strong>for</strong> the element of area in the xy-plane. By choosing thesubregions to be small rectangles each of area ∆A =∆x∆y, <strong>and</strong> letting both ∆x187


MULTIPLE INTEGRALSydSdxdA = dxdyRVdyUcTCa b xFigure 6.1 A simple curve C in the xy-plane, enclosing a region R.<strong>and</strong> ∆y → 0, we can also write the integral as∫∫I = f(x, y) dx dy, (6.2)Rwhere we have written out the element of area explicitly as the product of thetwo coordinate differentials (see figure 6.1).Some authors use a single integration symbol whatever the dimension of theintegral; others use as many symbols as the dimension. In different circumstancesboth have their advantages. We will adopt the convention used in (6.1) <strong>and</strong> (6.2),that as many integration symbols will be used as differentials explicitly written.The <strong>for</strong>m (6.2) gives us a clue as to how we may proceed in the evaluationof a double integral. Referring to figure 6.1, the limits on the integration maybewrittenasanequationc(x, y) = 0 giving the boundary curve C. However, anexplicit statement of the limits can be written in two distinct ways.One way of evaluating the integral is first to sum up the contributions fromthe small rectangular elemental areas in a horizontal strip of width dy (as shownin the figure) <strong>and</strong> then to combine the contributions of these horizontal strips tocover the region R. In this case, we write∫ y=d{∫ x=x2(y)}I =f(x, y) dx dy, (6.3)y=cx=x 1(y)where x = x 1 (y) <strong>and</strong>x = x 2 (y) are the equations of the curves TSV <strong>and</strong> TUVrespectively. This expression indicates that first f(x, y) istobeintegratedwithrespect to x (treating y as a constant) between the values x = x 1 (y) <strong>and</strong>x = x 2 (y)<strong>and</strong> then the result, considered as a function of y, is to be integrated between thelimits y = c <strong>and</strong> y = d. Thus the double integral is evaluated by expressing it interms of two single integrals called iterated (or repeated) integrals.188


6.1 DOUBLE INTEGRALSAn alternative way of evaluating the integral, however, is first to sum up thecontributions from the elemental rectangles arranged into vertical strips <strong>and</strong> thento combine these vertical strips to cover the region R. Wethenwrite∫ x=b{∫ y=y2(x)}I =f(x, y) dy dx, (6.4)x=ay=y 1(x)where y = y 1 (x) <strong>and</strong>y = y 2 (x) are the equations of the curves STU <strong>and</strong> SVUrespectively. In going to (6.4) from (6.3), we have essentially interchanged theorder of integration.In the discussion above we assumed that the curve C was such that any lineparallel to either the x- ory-axis intersected C at most twice. In general, providedf(x, y) is continuous everywhere in R <strong>and</strong> the boundary curve C has this simpleshape, the same result is obtained irrespective of the order of integration. In caseswhere the region R has a more complicated shape, it can usually be subdividedinto smaller simpler regions R 1 , R 2 etc. that satisfy this criterion. The doubleintegral over R is then merely the sum of the double integrals over the subregions.◮Evaluate the double integral∫∫I = x 2 ydxdy,Rwhere R is the triangular area bounded by the lines x =0, y =0<strong>and</strong> x + y =1. Reversethe order of integration <strong>and</strong> demonstrate that the same result is obtained.The area of integration is shown in figure 6.2. Suppose we choose to carry out theintegration with respect to y first. With x fixed, the range of y is 0 to 1 − x. Wecanthere<strong>for</strong>e write∫ x=1{∫ y=1−x}I =x 2 ydy dxx=0∫ x=1y=0] y=1−x[ x 2 y 2∫ 1x 2 (1 − x) 2=dx =dx = 1x=0 2y=00 2 60 .Alternatively, we may choose to per<strong>for</strong>m the integration with respect to x first. With yfixed, the range of x is 0 to 1 − y, so we have∫ y=1{∫ x=1−y}I =x 2 ydx dy=y=0∫ y=1y=0[ x 3 y3x=0] x=1−yx=0∫ 1dx =0(1 − y) 3 y3dy = 1 60 .As expected, we obtain the same result irrespective of the order of integration. ◭We may avoid the use of braces in expressions such as (6.3) <strong>and</strong> (6.4) by writing(6.4), <strong>for</strong> example, asI =∫ badx∫ y2(x)y 1(x)dy f(x, y),where it is understood that each integral symbol acts on everything to its right,189


MULTIPLE INTEGRALSy1dyRx + y =100dx1xFigure 6.2 The triangular region whose sides are the axes x =0,y =0<strong>and</strong>the line x + y =1.<strong>and</strong> that the order of integration is from right to left. So, in this example, theintegr<strong>and</strong> f(x, y) is first to be integrated with respect to y <strong>and</strong> then with respectto x. With the double integral expressed in this way, we will no longer write theindependent variables explicitly in the limits of integration, since the differentialof the variable with respect to which we are integrating is always adjacent to therelevant integral sign.Using the order of integration in (6.3), we could also write the double integral asI =∫ dcdy∫ x2(y)x 1(y)dx f(x, y).Occasionally, however, interchange of the order of integration in a double integralis not permissible, as it yields a different result. For example, difficulties mightarise if the region R were unbounded with some of the limits infinite, though inmany cases involving infinite limits the same result is obtained whichever orderof integration is used. Difficulties can also occur if the integr<strong>and</strong> f(x, y) has anydiscontinuities in the region R or on its boundary C.6.2 Triple integralsThe above discussion <strong>for</strong> double integrals can easily be extended to triple integrals.Consider the function f(x, y, z) defined in a closed three-dimensional region R.Proceeding as we did <strong>for</strong> double integrals, let us divide the region R into Nsubregions ∆R p of volume ∆V p , p =1, 2,...,N,<strong>and</strong>let(x p ,y p ,z p ) be any point inthe subregion ∆R p . Now we <strong>for</strong>m the sumS =N∑f(x p ,y p ,z p )∆V p ,p=1190


6.3 APPLICATIONS OF MULTIPLE INTEGRALS<strong>and</strong> let N →∞as each of the volumes ∆V p → 0. If the sum S tends to a uniquelimit, I, then this is called the triple integral of f(x, y, z) over the region R <strong>and</strong> iswritten∫I = f(x, y, z) dV , (6.5)Rwhere dV st<strong>and</strong>s <strong>for</strong> the element of volume. By choosing the subregions to besmall cuboids, each of volume ∆V =∆x∆y∆z, <strong>and</strong> proceeding to the limit, wecanalsowritetheintegralas∫∫∫I = f(x, y, z) dx dy dz, (6.6)Rwhere we have written out the element of volume explicitly as the product of thethree coordinate differentials. Extending the notation used <strong>for</strong> double integrals,we may write triple integrals as three iterated integrals, <strong>for</strong> example,I =∫ x2x 1dx∫ y2(x)y 1(x)dy∫ z2(x,y)z 1(x,y)dz f(x, y, z),where the limits on each of the integrals describe the values that x, y <strong>and</strong> z takeon the boundary of the region R. As <strong>for</strong> double integrals, in most cases the orderof integration does not affect the value of the integral.We can extend these ideas to define multiple integrals of higher dimensionalityin a similar way.6.3 Applications of multiple integralsMultiple integrals have many uses in the physical sciences, since there are numerousphysical quantities which can be written in terms of them. We now discuss afew of the more common examples.6.3.1 Areas <strong>and</strong> volumesMultiple integrals are often used in finding areas <strong>and</strong> volumes. For example, theintegral∫ ∫∫A = dA = dx dyRRis simply equal to the area of the region R. Similarly, if we consider the surfacez = f(x, y) in three-dimensional Cartesian coordinates then the volume under thissurface that st<strong>and</strong>s vertically above the region R is given by the integral∫ ∫∫V = zdA= f(x, y) dx dy,Rwhere volumes above the xy-plane are counted as positive, <strong>and</strong> those below asnegative.191R


MULTIPLE INTEGRALSzcdV = dx dy dzdzdxdybyaxFigure 6.3 The tetrahedron bounded by the coordinate surfaces <strong>and</strong> theplane x/a + y/b + z/c = 1 is divided up into vertical slabs, the slabs intocolumns <strong>and</strong> the columns into small boxes.◮Find the volume of the tetrahedron bounded by the three coordinate surfaces x =0, y =0<strong>and</strong> z =0<strong>and</strong> the plane x/a + y/b + z/c =1.Referring to figure 6.3, the elemental volume of the shaded region is given by dV = zdxdy,<strong>and</strong> we must integrate over the triangular region R in the xy-plane whose sides are x =0,y =0<strong>and</strong>y = b − bx/a. The total volume of the tetrahedron is there<strong>for</strong>e given by∫∫V = zdxdy=R= c= c∫ adx0∫ a0∫ a0∫ b−bx/a0(dy c 1 − y b − x )adx[y − y22b − xy a( bx2dx2a − bx 2 a + b 2] y=b−bx/ay=0)= abc6 . ◭Alternatively, we can write the volume of a three-dimensional region R as∫ ∫∫∫V = dV = dx dy dz, (6.7)RRwhere the only difficulty occurs in setting the correct limits on each of theintegrals. For the above example, writing the volume in this way corresponds todividing the tetrahedron into elemental boxes of volume dx dy dz (as shown infigure 6.3); integration over z then adds up the boxes to <strong>for</strong>m the shaded columnin the figure. The limits of integration are z =0toz = c ( 1 − y/b − x/a ) ,<strong>and</strong>192


6.3 APPLICATIONS OF MULTIPLE INTEGRALSthe total volume of the tetrahedron is given byV =∫ a0dx∫ b−bx/a0dy∫ c(1−y/b−x/a)0dz, (6.8)which clearly gives the same result as above. This method is illustrated further inthe following example.◮Find the volume of the region bounded by the paraboloid z = x 2 + y 2z =2y.<strong>and</strong> the planeThe required region is shown in figure 6.4. In order to write the volume of the region inthe <strong>for</strong>m (6.7), we must deduce the limits on each of the integrals. Since the integrationscan be per<strong>for</strong>med in any order, let us first divide the region into vertical slabs of thicknessdy perpendicular to the y-axis, <strong>and</strong> then as shown in the figure we cut each slab intohorizontal strips of height dz, <strong>and</strong> each strip into elemental boxes of volume dV = dx dy dz.Integrating first with respect to x (adding up the elemental boxes to get a horizontal strip),the limits on x are x = − √ z − y 2 to x = √ z − y 2 . Now integrating with respect to z(adding up the strips to <strong>for</strong>m a vertical slab) the limits on z are z = y 2 to z =2y. Finally,integrating with respect to y (adding up the slabs to obtain the required region), the limitson y are y =0<strong>and</strong>y = 2, the solutions of the simultaneous equations z =0 2 + y 2 <strong>and</strong>z =2y. So the volume of the region isV ==∫ 20∫ 20dy∫ 2yy 2∫ √ z−y 2dz dx =−√z−y 2dy [ 43 (z − y2 ) 3/2] z=2yz=y 2 =∫ 20∫ 20∫ 2ydy dz 2 √ z − y 2y 2dy 4 3 (2y − y2 ) 3/2 .The integral over y may be evaluated straight<strong>for</strong>wardly by making the substitution y =1+sinu, <strong>and</strong> gives V = π/2. ◭In general, when calculating the volume (area) of a region, the volume (area)elements need not be small boxes as in the previous example, but may be of anyconvenient shape. The latter is usually chosen to make evaluation of the integralas simple as possible.6.3.2 Masses, centres of mass <strong>and</strong> centroidsIt is sometimes necessary to calculate the mass of a given object having a nonuni<strong>for</strong>mdensity. Symbolically, this mass is given simply by∫M = dM,where dM is the element of mass <strong>and</strong> the integral is taken over the extent of theobject. For a solid three-dimensional body the element of mass is just dM = ρdV,where dV is an element of volume <strong>and</strong> ρ is the variable density. For a laminarbody (i.e. a uni<strong>for</strong>m sheet of material) the element of mass is dM = σdA,whereσ is the mass per unit area of the body <strong>and</strong> dA is an area element. Finally, <strong>for</strong>a body in the <strong>for</strong>m of a thin wire we have dM = λds,whereλ is the mass per193


MULTIPLE INTEGRALSzz =2yz = x 2 + y 20 2 ydV = dx dy dzxFigure 6.4 The region bounded by the paraboloid z = x 2 + y 2 <strong>and</strong> the planez =2y is divided into vertical slabs, the slabs into horizontal strips <strong>and</strong> thestrips into boxes.unit length <strong>and</strong> ds is an element of arc length along the wire. When evaluatingthe required integral, we are free to divide up the body into mass elements inthe most convenient way, provided that over each mass element the density isapproximately constant.◮ Find the mass of the tetrahedron bounded by the three coordinate surfaces <strong>and</strong> the planex/a + y/b + z/c =1, if its density is given by ρ(x, y, z) =ρ 0 (1 + x/a).From (6.8), we can immediately write down the mass of the tetrahedron as∫ (M = ρ 0 1+ x ) ∫ a (dV = dx ρ 0 1+ x ) ∫ b−bx/a ∫ c(1−y/b−x/a)dydz,R a0a 00where we have taken the density outside the integrations with respect to z <strong>and</strong> y since itdepends only on x. There<strong>for</strong>e the integrations with respect to z <strong>and</strong> y proceed exactly asthey did when finding the volume of the tetrahedron, <strong>and</strong> we have∫ a (M = cρ 0 dx 1+ x ) ( bx 2a 2a − bx 2 a + b 20). (6.9)We could have arrived at (6.9) more directly by dividing the tetrahedron into triangularslabs of thickness dx perpendicular to the x-axis (see figure 6.3), each of which is ofconstant density, since ρ depends on x alone. A slab at a position x has volume dV =1c(1 − x/a)(b − bx/a) dx <strong>and</strong> mass dM = ρdV = ρ 2 0(1 + x/a) dV . Integrating over x weagain obtain (6.9). This integral is easily evaluated <strong>and</strong> gives M = 5 abcρ 24 0. ◭194


6.3 APPLICATIONS OF MULTIPLE INTEGRALSThe coordinates of the centre of mass of a solid or laminar body may also bewritten as multiple integrals. The centre of mass of a body has coordinates ¯x, ȳ,¯z given by the three equations∫¯x∫∫dM =∫xdMȳ dM = ydM∫ ∫¯z dM = zdM,where again dM is an element of mass as described above, x, y, z are thecoordinates of the centre of mass of the element dM <strong>and</strong> the integrals are takenover the entire body. Obviously, <strong>for</strong> any body that lies entirely in, or is symmetricalabout, the xy-plane (say), we immediately have ¯z = 0. For completeness, we notethat the three equations above can be written as the single vector equation (seechapter 7)¯r = 1 ∫r dM,Mwhere ¯r is the position vector of the body’s centre of mass with respect to theorigin, r is the position vector of the centre of mass of the element dM <strong>and</strong>M = ∫ dM is the total mass of the body. As previously, we may divide the bodyinto the most convenient mass elements <strong>for</strong> evaluating the necessary integrals,provided each mass element is of constant density.We further note that the coordinates of the centroid of a body are defined asthose of its centre of mass if the body had uni<strong>for</strong>m density.◮Find the centre of mass of the solid hemisphere bounded by the surfaces x 2 + y 2 + z 2 = a 2<strong>and</strong> the xy-plane, assuming that it has a uni<strong>for</strong>m density ρ.Referring to figure 6.5, we know from symmetry that the centre of mass must lie onthe z-axis. Let us divide the hemisphere into volume elements that are circular slabs ofthickness dz parallel to the xy-plane. For a slab at a height z, the mass of the element isdM = ρdV = ρπ(a 2 − z 2 ) dz. Integrating over z, we find that the z-coordinate of the centreof mass of the hemisphere is given by¯z∫ a0ρπ(a 2 − z 2 ) dz =∫ a0zρπ(a 2 − z 2 ) dz.The integrals are easily evaluated <strong>and</strong> give ¯z =3a/8. Since the hemisphere is of uni<strong>for</strong>mdensity, this is also the position of its centroid. ◭6.3.3 Pappus’ theoremsThe theorems of Pappus (which are about seventeen centuries old) relate centroidsto volumes of revolution <strong>and</strong> areas of surfaces, discussed in chapter 2, <strong>and</strong> may beuseful <strong>for</strong> finding one quantity given another that can be calculated more easily.195


MULTIPLE INTEGRALSza√a2 − z 2dzxaayFigure 6.5 The solid hemisphere bounded by the surfaces x 2 + y 2 + z 2 = a 2<strong>and</strong> the xy-plane.ydAAyȳxFigure 6.6 An area A in the xy-plane, which may be rotated about the x-axisto <strong>for</strong>m a volume of revolution.If a plane area is rotated about an axis that does not intersect it then the solidso generated is called a volume of revolution. Pappus’ first theorem states that thevolume of such a solid is given by the plane area A multiplied by the distancemoved by its centroid (see figure 6.6). This may be proved by considering thedefinition of the centroid of the plane area as the position of the centre of massif the density is uni<strong>for</strong>m, so thatȳ = 1 ∫ydA.ANow the volume generated by rotating the plane area about the x-axis is given by∫V = 2πy dA =2πȳA,which is the area multiplied by the distance moved by the centroid.196


6.3 APPLICATIONS OF MULTIPLE INTEGRALSydsyȳxFigure 6.7 A curve in the xy-plane, which may be rotated about the x-axisto <strong>for</strong>m a surface of revolution.Pappus’ second theorem states that if a plane curve is rotated about a coplanaraxis that does not intersect it then the area of the surface of revolution so generatedis given by the length of the curve L multiplied by the distance moved by itscentroid (see figure 6.7). This may be proved in a similar manner to the firsttheorem by considering the definition of the centroid of a plane curve,ȳ = 1 L∫yds,<strong>and</strong> noting that the surface area generated is given by∫S = 2πy ds =2πȳL,which is equal to the length of the curve multiplied by the distance moved by itscentroid.◮ A semicircular uni<strong>for</strong>m lamina is freely suspended from one of its corners. Show that itsstraight edge makes an angle of 23.0 ◦ with the vertical.Referring to figure 6.8, the suspended lamina will have its centre of gravity C verticallybelow the suspension point <strong>and</strong> its straight edge will make an angle θ =tan −1 (d/a) withthe vertical, where 2a is the diameter of the semicircle <strong>and</strong> d is the distance of its centreof mass from the diameter.Since rotating the lamina about the diameter generates a sphere of volume 4 3 πa3 , Pappus’first theorem requires that43 πa3 =2πd × 1 2 πa2 .Hence d = 4a4<strong>and</strong> θ =tan−13π 3π =23.0◦ . ◭197


MULTIPLE INTEGRALSaθdCFigure 6.8Suspending a semicircular lamina from one of its corners.6.3.4 Moments of inertiaFor problems in rotational mechanics it is often necessary to calculate the momentof inertia of a body about a given axis. This is defined by the multiple integral∫I = l 2 dM,where l is the distance of a mass element dM from the axis. We may again choosemass elements convenient <strong>for</strong> evaluating the integral. In this case, however, inaddition to elements of constant density we require all parts of each element tobe at approximately the same distance from the axis about which the moment ofinertia is required.◮ Find the moment of inertia of a uni<strong>for</strong>m rectangular lamina of mass M with sides a <strong>and</strong>b about one of the sides of length b.Referring to figure 6.9, we wish to calculate the moment of inertia about the y-axis.We there<strong>for</strong>e divide the rectangular lamina into elemental strips parallel to the y-axis ofwidth dx. The mass of such a strip is dM = σb dx, whereσ is the mass per unit area ofthe lamina. The moment of inertia of a strip at a distance x from the y-axisissimplydI = x 2 dM = σbx 2 dx. The total moment of inertia of the lamina about the y-axis isthere<strong>for</strong>eI =∫ a0σbx 2 dx = σba33 .Since the total mass of the lamina is M = σab, wecanwriteI = 1 3 Ma2 . ◭198


6.4 CHANGE OF VARIABLES IN MULTIPLE INTEGRALSybdM = σb dxdxaxFigure 6.9 A uni<strong>for</strong>m rectangular lamina of mass M with sides a <strong>and</strong> b canbe divided into vertical strips.6.3.5 Mean values of functionsIn chapter 2 we discussed average values <strong>for</strong> functions of a single variable. Thisis easily extended to functions of several variables. Let us consider, <strong>for</strong> example,a function f(x, y) defined in some region R of the xy-plane. Then the averagevalue ¯f of the function is given by∫ ∫¯f dA = f(x, y) dA. (6.10)RThis definition is easily extended to three (<strong>and</strong> higher) dimensions; if a functionf(x, y, z) is defined in some three-dimensional region of space R then the averagevalue ¯f of the function is given by∫ ∫¯f dV = f(x, y, z) dV . (6.11)RRR◮A tetrahedron is bounded by the three coordinate surfaces <strong>and</strong> the plane x/a+y/b+z/c =1 <strong>and</strong> has density ρ(x, y, z) =ρ 0 (1 + x/a). Find the average value of the density.From (6.11), the average value of the density is given by∫ ∫¯ρ dV = ρ(x, y, z) dV .RRNow the integral on the LHS is just the volume of the tetrahedron, which we found insubsection 6.3.1 to be V = 1 5abc, <strong>and</strong> the integral on the RHS is its mass M = abcρ 6 24 0,calculated in subsection 6.3.2. There<strong>for</strong>e ¯ρ = M/V = 5 ρ 4 0. ◭6.4 Change of variables in multiple integralsIt often happens that, either because of the <strong>for</strong>m of the integr<strong>and</strong> involved orbecause of the boundary shape of the region of integration, it is desirable to199


MULTIPLE INTEGRALSyu =constantv =constantNKMLRCFigure 6.10 A region of integration R overlaid with a grid <strong>for</strong>med by thefamily of curves u =constant<strong>and</strong>v = constant. The parallelogram KLMNdefines the area element dA uv .xexpress a multiple integral in terms of a new set of variables. We now considerhow to do this.6.4.1 Change of variables in double integralsLet us begin by examining the change of variables in a double integral. Supposethat we require to change an integral∫∫I = f(x, y) dx dy,Rin terms of coordinates x <strong>and</strong> y, into one expressed in new coordinates u <strong>and</strong> v,given in terms of x <strong>and</strong> y by differentiable equations u = u(x, y) <strong>and</strong>v = v(x, y)with inverses x = x(u, v) <strong>and</strong>y = y(u, v). The region R in the xy-plane <strong>and</strong> thecurve C that bounds it will become a new region R ′ <strong>and</strong> a new boundary C ′ inthe uv-plane, <strong>and</strong> so we must change the limits of integration accordingly. Also,the function f(x, y) becomes a new function g(u, v) of the new coordinates.Now the part of the integral that requires most consideration is the area element.In the xy-plane the element is the rectangular area dA xy = dx dy generated byconstructing a grid of straight lines parallel to the x- <strong>and</strong>y- axes respectively.Our task is to determine the corresponding area element in the uv-coordinates. Ingeneral the corresponding element dA uv will not be the same shape as dA xy , butthis does not matter since all elements are infinitesimally small <strong>and</strong> the value ofthe integr<strong>and</strong> is considered constant over them. Since the sides of the area elementare infinitesimal, dA uv will in general have the shape of a parallelogram. We canfind the connection between dA xy <strong>and</strong> dA uv by considering the grid <strong>for</strong>med by thefamily of curves u =constant<strong>and</strong>v = constant, as shown in figure 6.10. Since v200


6.4 CHANGE OF VARIABLES IN MULTIPLE INTEGRALSis constant along the line element KL, the latter has components (∂x/∂u) du <strong>and</strong>(∂y/∂u) du in the directions of the x- <strong>and</strong>y-axes respectively. Similarly, since uis constant along the line element KN, the latter has corresponding components(∂x/∂v) dv <strong>and</strong> (∂y/∂v) dv. Using the result <strong>for</strong> the area of a parallelogram givenin chapter 7, we find that the area of the parallelogram KLMN is given by∣ dA uv =∂x∣ ∂u du∂y ∂x ∂y ∣∣∣dv − dv∂v ∂v ∂u du =∂x ∂y∣ ∂u ∂v − ∂x∂y∂v ∂u∣ du dv.Defining the Jacobian of x, y with respect to u, v as∂(x, y)J =∂(u, v) ≡ ∂x ∂y∂u ∂v − ∂x ∂y∂v ∂u ,we havedA uv =∂(x, y)∣ ∂(u, v) ∣ du dv.The reader acquainted with determinants will notice that the Jacobian can alsobe written as the 2 × 2 determinant∣ ∣∣∣∣∣∣∣ ∂x ∂y∂(x, y)J =∂(u, v) = ∂u ∂u∂x ∂y.∣∂v ∂vSuch determinants can be evaluated using the methods of chapter 8.So, in summary, the relationship between the size of the area element generatedby dx, dy <strong>and</strong> the size of the corresponding area element generated by du, dv isdx dy =∂(x, y)∣ ∂(u, v) ∣ du dv.This equality should be taken as meaning that when trans<strong>for</strong>ming from coordinatesx, y to coordinates u, v, the area element dx dy should be replaced by theexpression on the RHS of the above equality. Of course, the Jacobian can, <strong>and</strong>in general will, vary over the region of integration. We may express the doubleintegral in either coordinate system as∫∫∫∫I = f(x, y) dx dy = g(u, v)∂(x, y)∣RR ∂(u, v) ∣ du dv. (6.12)′When evaluating the integral in the new coordinate system, it is usually advisableto sketch the region of integration R ′ in the uv-plane.201


MULTIPLE INTEGRALS◮Evaluate the double integral∫∫I =(a + √ )x 2 + y 2 dx dy,where R is the region bounded by the circle x 2 + y 2 = a 2 .00RIn Cartesian coordinates, the integral may be written∫ a ∫ √ a 2 −x 2I = dx−a − √ dy(a + √ )x 2 + y 2 ,a 2 −x 2<strong>and</strong> can be calculated directly. However, because of the circular boundary of the integrationregion, a change of variables to plane polar coordinates ρ, φ is indicated. The relationshipbetween Cartesian <strong>and</strong> plane polar coordinates is given by x = ρ cos φ <strong>and</strong> y = ρ sin φ.Using (6.12) we can there<strong>for</strong>e write∫∫I = (a + ρ)∂(x, y)∣R ′ ∂(ρ, φ) ∣ dρ dφ,where R ′ is the rectangular region in the ρφ-plane whose sides are ρ =0,ρ = a, φ =0<strong>and</strong> φ =2π. The Jacobian is easily calculated, <strong>and</strong> we obtain∣ ∂(x, y) ∣∣∣J =∂(ρ, φ) = cos φ sin φ−ρ sin φ ρcos φ ∣ = ρ(cos2 φ +sin 2 φ)=ρ.So the relationship between the area elements in Cartesian <strong>and</strong> in plane polar coordinates isdx dy = ρdρdφ.There<strong>for</strong>e, when expressed in plane polar coordinates, the integral is given by∫∫I = (a + ρ)ρdρdφR∫ ′ 2π ∫ a[ ] aρ2 a= dφ dρ (a + ρ)ρ =2π2 + ρ3= 5πa3303 . ◭6.4.2 Evaluation of the integral I = ∫ ∞−∞ e−x2 dxBy making a judicious change of variables, it is sometimes possible to evaluatean integral that would be intractable otherwise. An important example of thismethod is provided by the evaluation of the integralI =∫ ∞−∞e −x2 dx.Its value may be found by first constructing I 2 , as follows:I 2 =∫ ∞−∞e −x2 dx∫ ∞−∞∫ ∞∫ ∞e −y2 dy = dx dy e −(x2 +y 2 )−∞ −∞∫∫= e −(x2 +y 2) dx dy,202R


6.4 CHANGE OF VARIABLES IN MULTIPLE INTEGRALSya−aax−aFigure 6.11 The regions used to illustrate the convergence properties of theintegral I(a) = ∫ a−a e−x2 dx as a →∞.where the region R is the whole xy-plane. Then, trans<strong>for</strong>ming to plane polarcoordinates, we find∫∫∫ 2π ∫ ∞I 2 = e −ρ2 ρdρdφ= dφ dρ ρe −ρ2 =2π[− 1 2 e−ρ2] ∞= π.R ′ 0 00There<strong>for</strong>e the original integral is given by I = √ π. Because the integr<strong>and</strong> is aneven function of x, it follows that the value of the integral from 0 to ∞ is simply√ π/2.We note, however, that unlike in all the previous examples, the regions ofintegration R <strong>and</strong> R ′ are both infinite in extent (i.e. unbounded). It is there<strong>for</strong>eprudent to derive this result more rigorously; this we do by considering theintegral∫ aI(a) = e −x2 dx.−aWe then have∫∫I 2 (a) = e −(x2 +y 2) dx dy,Rwhere R is the square of side 2a centred on the origin. Referring to figure 6.11,since the integr<strong>and</strong> is always positive the value of the integral taken over thesquare lies between the value of the integral taken over the region bounded bythe inner circle of radius a <strong>and</strong> the value of the integral taken over the outercircle of radius √ 2a. Trans<strong>for</strong>ming to plane polar coordinates as above, we may203


MULTIPLE INTEGRALSzRTu = c 1v = c 2SPQw = c 3CyxFigure 6.12 A three-dimensional region of integration R, showing an elementof volume in u, v, w coordinates <strong>for</strong>med by the coordinate surfacesu =constant,v =constant,w =constant.evaluate the integrals over the inner <strong>and</strong> outer circles respectively, <strong>and</strong> we find(π 1 − e −a2) (


6.4 CHANGE OF VARIABLES IN MULTIPLE INTEGRALSw are constant, <strong>and</strong> so PQ has components (∂x/∂u) du, (∂y/∂u) du <strong>and</strong> (∂z/∂u) duin the directions of the x-, y- <strong>and</strong>z- axes respectively. The components of theline elements PS <strong>and</strong> ST are found by replacing u by v <strong>and</strong> w respectively.The expression <strong>for</strong> the volume of a parallelepiped in terms of the componentsof its edges with respect to the x-, y- <strong>and</strong>z-axes is given in chapter 7. Using this,we find that the element of volume in u, v, w coordinates is given bydV uvw =∂(x, y, z)∣∂(u, v, w) ∣ du dv dw,where the Jacobian of x, y, z with respect to u, v, w is a short-h<strong>and</strong> <strong>for</strong> a 3 × 3determinant:∣ ∣∣∣∣∣∣∣∣∣∣∣ ∂x ∂y ∂z∂u ∂u ∂u∂(x, y, z)∂(u, v, w) ≡ ∂x ∂y ∂z∂v ∂v ∂v.∣∂x∂wSo, in summary, the relationship between the elemental volumes in multipleintegrals <strong>for</strong>mulated in the two coordinate systems is given in Jacobian <strong>for</strong>m bydx dy dz =∂(x, y, z)∣∂(u, v, w) ∣ du dv dw,<strong>and</strong> we can write a triple integral in either set of coordinates as∫∫∫∫∫∫I = f(x, y, z) dx dy dz = g(u, v, w)∂(x, y, z)∣RR ∂(u, v, w) ∣ du dv dw.′∂y∂w∂z∂w∣◮ Find an expression <strong>for</strong> a volume element in spherical polar coordinates, <strong>and</strong> hence calculatethe moment of inertia about a diameter of a uni<strong>for</strong>m sphere of radius a <strong>and</strong> mass M.Spherical polar coordinates r, θ, φ are defined byx = r sin θ cos φ, y = r sin θ sin φ, z = r cos θ(<strong>and</strong> are discussed fully in chapter 10). The required Jacobian is there<strong>for</strong>e∣ ∣∣∣∣∣∂(x, y, z)sin θ cos φ sin θ sin φ cos θJ =∂(r, θ, φ) = r cos θ cos φ rcos θ sin φ −r sin θ−r sin θ sin φ rsin θ cos φ 0 ∣ .The determinant is most easily evaluated by exp<strong>and</strong>ing it with respect to the last column(see chapter 8), which givesJ =cosθ(r 2 sin θ cos θ)+r sin θ(r sin 2 θ)= r 2 sin θ(cos 2 θ +sin 2 θ)=r 2 sin θ.There<strong>for</strong>e the volume element in spherical polar coordinates is given bydV =∂(x, y, z)∂(r, θ, φ) dr dθ dφ = r2 sin θdrdθdφ,205


MULTIPLE INTEGRALSwhich agrees with the result given in chapter 10.If we place the sphere with its centre at the origin of an x, y, z coordinate system thenits moment of inertia about the z-axis (which is, of course, a diameter of the sphere) isI =∫ (x 2 + y 2) dM = ρ∫ (x 2 + y 2) dV ,where the integral is taken over the sphere, <strong>and</strong> ρ is the density. Using spherical polarcoordinates, we can write this as∫∫∫(I = ρ r 2 sin 2 θ ) r 2 sin θdrdθdφ= ρV∫ 2π0dφ∫ π0dθ sin 3 θ= ρ × 2π × 4 × 1 3 5 a5 = 8 15 πa5 ρ.Since the mass of the sphere is M = 4 3 πa3 ρ, the moment of inertia can also be written asI = 2 5 Ma2 . ◭∫ a0dr r 46.4.4 General properties of JacobiansAlthough we will not prove it, the general result <strong>for</strong> a change of coordinates inan n-dimensional integral from a set x i to a set y j (where i <strong>and</strong> j both run from1ton) isdx 1 dx 2 ···dx n =∂(x 1 ,x 2 ,...,x n )∣ ∂(y 1 ,y 2 ,...,y n ) ∣ dy 1 dy 2 ···dy n ,where the n-dimensional Jacobian can be written as an n × n determinant (seechapter 8) in an analogous way to the two- <strong>and</strong> three-dimensional cases.For readers who already have sufficient familiarity with matrices (see chapter 8)<strong>and</strong> their properties, a fairly compact proof of some useful general propertiesof Jacobians can be given as follows. Other readers should turn straight to theresults (6.16) <strong>and</strong> (6.17) <strong>and</strong> return to the proof at some later time.Consider three sets of variables x i , y i <strong>and</strong> z i , with i running from 1 to n <strong>for</strong>each set. From the chain rule in partial differentiation (see (5.17)), we know that∂x in∑ ∂x i ∂y k=. (6.13)∂z j ∂y k ∂z jk=1Now let A, B <strong>and</strong> C be the matrices whose ijth elements are ∂x i /∂y j , ∂y i /∂z j <strong>and</strong>∂x i /∂z j respectively. We can then write (6.13) as the matrix productn∑c ij = a ik b kj or C = AB. (6.14)k=1We may now use the general result <strong>for</strong> the determinant of the product of twomatrices, namely |AB| = |A||B|, <strong>and</strong> recall that the JacobianJ xy = ∂(x 1,...,x n )= |A|, (6.15)∂(y 1 ,...,y n )206


6.5 EXERCISES<strong>and</strong> similarly <strong>for</strong> J yz <strong>and</strong> J xz . On taking the determinant of (6.14), we there<strong>for</strong>eobtainJ xz = J xy J yzor, in the usual notation,∂(x 1 ,...,x n )∂(z 1 ,...,z n ) = ∂(x 1,...,x n ) ∂(y 1 ,...,y n )∂(y 1 ,...,y n ) ∂(z 1 ,...,z n ) . (6.16)As a special case, if the set z i istakentobeidenticaltothesetx i ,<strong>and</strong>theobvious result J xx = 1 is used, we obtainJ xy J yx =1or, in the usual notation,[ ] −1∂(x 1 ,...,x n )∂(y 1 ,...,y n ) = ∂(y1 ,...,y n ). (6.17)∂(x 1 ,...,x n )The similarity between the properties of Jacobians <strong>and</strong> those of derivatives isapparent, <strong>and</strong> to some extent is suggested by the notation. We further note from(6.15) that since |A| = |A T |,whereA T is the transpose of A, we can interchange therows <strong>and</strong> columns in the determinantal <strong>for</strong>m of the Jacobian without changingits value.6.5 Exercises6.1 Identify the curved wedge bounded by the surfaces y 2 =4ax, x + z = a <strong>and</strong>z = 0, <strong>and</strong> hence calculate its volume V .6.2 Evaluate the volume integral of x 2 + y 2 + z 2 over the rectangular parallelepipedbounded by the six surfaces x = ±a, y = ±b <strong>and</strong> z = ±c.6.3 Find the volume integral of x 2 y over the tetrahedral volume bounded by theplanes x =0,y =0,z =0,<strong>and</strong>x + y + z =1.6.4 Evaluate the surface integral of f(x, y) over the rectangle 0 ≤ x ≤ a, 0≤ y ≤ b<strong>for</strong> the functionsx(a) f(x, y) =x 2 + y , (b) f(x, y) =(b − y + 2 x)−3/2 .6.5 Calculate the volume of an ellipsoid as follows:(a) Prove that the area of the ellipsex 2a + y22 b =1 2is πab.(b) Use this result to obtain an expression <strong>for</strong> the volume of a slice of thicknessdz of the ellipsoidx 2a + y22 b + z22 c =1. 2Hence show that the volume of the ellipsoid is 4πabc/3.207


MULTIPLE INTEGRALS6.6 The function(Ψ(r) =A 2 − Zr )e −Zr/2aagives the <strong>for</strong>m of the quantum-mechanical wavefunction representing the electronin a hydrogen-like atom of atomic number Z, when the electron is in its firstallowed spherically symmetric excited state. Here r is the usual spherical polarcoordinate, but, because of the spherical symmetry, the coordinates θ <strong>and</strong> φ donot appear explicitly in Ψ. Determine the value that A (assumed real) must haveif the wavefunction is to be correctly normalised, i.e. if the volume integral of|Ψ| 2 over all space is to be equal to unity.6.7 In quantum mechanics the electron in a hydrogen atom in some particular stateis described by a wavefunction Ψ, which is such that |Ψ| 2 dV is the probability offinding the electron in the infinitesimal volume dV . In spherical polar coordinatesΨ=Ψ(r, θ, φ) <strong>and</strong>dV = r 2 sin θdrdθdφ. Two such states are described by( ) 1/2 ( ) 3/2 1 1Ψ 1 =2e −r/a 0,4π a 0( ) 1/2 ( ) 3/2 31Ψ 2 = − sin θe iφ re −r/2a 0√ .8π2a 0 a 0 3(a) Show that each Ψ i is normalised, i.e. the integral over all space ∫ |Ψ| 2 dV isequal to unity – physically, this means that the electron must be somewhere.(b) The (so-called) dipole matrix element between the states 1 <strong>and</strong> 2 is given bythe integral∫p x = Ψ ∗ 1qr sin θ cos φ Ψ 2 dV ,where q is the charge on the electron. Prove that p x has the value −2 7 qa 0 /3 5 .6.8 A planar figure is <strong>for</strong>med from uni<strong>for</strong>m wire <strong>and</strong> consists of two equal semicirculararcs, each with its own closing diameter, joined so as to <strong>for</strong>m a letter ‘B’. Thefigure is freely suspended from its top left-h<strong>and</strong> corner. Show that the straightedge of the figure makes an angle θ with the vertical given by tan θ =(2+π) −1 .6.9 A certain torus has a circular vertical cross-section of radius a centredonahorizontal circle of radius c (> a).(a) Find the volume V <strong>and</strong> surface area A of the torus, <strong>and</strong> show that they canbe written asV = π24 (r2 o − ri 2 )(r o − r i ), A = π 2 (ro 2 − ri 2 ),where r o <strong>and</strong> r i are, respectively, the outer <strong>and</strong> inner radii of the torus.(b) Show that a vertical circular cylinder of radius c, coaxial with the torus,divides A in the ratioπc +2a : πc − 2a.6.10 A thin uni<strong>for</strong>m circular disc has mass M <strong>and</strong> radius a.(a) Prove that its moment of inertia about an axis perpendicular to its plane<strong>and</strong> passing through its centre is 1 2 Ma2 .(b) Prove that the moment of inertia of the same disc about a diameter is 1 4 Ma2 .208


6.5 EXERCISESThis is an example of the general result <strong>for</strong> planar bodies that the moment ofinertia of the body about an axis perpendicular to the plane is equal to the sumof the moments of inertia about two perpendicular axes lying in the plane; in anobvious notation∫ ∫I z = r 2 dm =∫(x 2 + y 2 ) dm =∫x 2 dm +y 2 dm = I y + I x .6.11 In some applications in mechanics the moment of inertia of a body about asingle point (as opposed to about an axis) is needed. The moment of inertia, I,about the origin of a uni<strong>for</strong>m solid body of density ρ is given by the volumeintegral∫I = (x 2 + y 2 + z 2 )ρdV.VShow that the moment of inertia of a right circular cylinder of radius a, length2b <strong>and</strong> mass M about its centre is( a2M2 + b236.12 The shape of an axially symmetric hard-boiled egg, of uni<strong>for</strong>m density ρ 0 ,isgiven in spherical polar coordinates by r = a(2 − cos θ), where θ is measuredfrom the axis of symmetry.(a) Prove that the mass M of the egg is M = 40 πρ 3 0a 3 .(b) Prove that the egg’s moment of inertia about its axis of symmetry is 342175 Ma2 .6.13 In spherical polar coordinates r, θ, φ the element of volume <strong>for</strong> a body thatis symmetrical about the polar axis is dV =2πr 2 sin θdrdθ, whilst its elementof surface area is 2πr sin θ[(dr) 2 + r 2 (dθ) 2 ] 1/2 . A particular surface is defined byr =2a cos θ, wherea is a constant <strong>and</strong> 0 ≤ θ ≤ π/2. Find its total surface area<strong>and</strong> the volume it encloses, <strong>and</strong> hence identify the surface.6.14 By expressing both the integr<strong>and</strong> <strong>and</strong> the surface element in spherical polarcoordinates, show that the surface integral∫x 2x 2 + y dS 2over the surface x 2 + y 2 = z 2 ,0≤ z ≤ 1, has the value π/ √ 2.6.15 By trans<strong>for</strong>ming to cylindrical polar coordinates, evaluate the integral∫ ∫ ∫I = ln(x 2 + y 2 ) dx dy dzover the interior of the conical region x 2 + y 2 ≤ z 2 ,0≤ z ≤ 1.6.16 Sketch the two families of curvesy 2 =4u(u − x), y 2 =4v(v + x),where u <strong>and</strong> v are parameters.By trans<strong>for</strong>ming to the uv-plane, evaluate the integral of y/(x 2 + y 2 ) 1/2 overthe part of the quadrant x>0, y>0 that is bounded by the lines x =0,y =0<strong>and</strong> the curve y 2 =4a(a − x).6.17 By making two successive simple changes of variables, evaluate∫ ∫ ∫I = x 2 dx dy dz).209


MULTIPLE INTEGRALSover the ellipsoidal regionx 2a + y22 b + z22 c ≤ 1. 26.18 Sketch the domain of integration <strong>for</strong> the integralI =∫ 1 ∫ 1/y0x=yy 3x exp[y2 (x 2 + x −2 )] dx dy<strong>and</strong> characterise its boundaries in terms of new variables u = xy <strong>and</strong> v = y/x.Show that the Jacobian <strong>for</strong> the change from (x, y) to(u, v) isequalto(2v) −1 ,<strong>and</strong>hence evaluate I.6.19 Sketch the part of the region 0 ≤ x, 0≤ y ≤ π/2 that is bounded by the curvesx =0,y =0,sinhx cos y =1<strong>and</strong>coshx sin y = 1. By making a suitable changeof variables, evaluate the integral∫ ∫I = (sinh 2 x +cos 2 y)sinh2x sin 2ydxdyover the bounded subregion.6.20 Define a coordinate system u, v whose origin coincides with that of the usualx, y system <strong>and</strong> whose u-axis coincides with the x-axis, whilst the v-axis makesan angle α with it. By considering the integral I = ∫ exp(−r 2 ) dA, wherer is theradial distance from the origin, over the area defined by 0 ≤ u


6.6 HINTS AND ANSWERS(a) Let R be a real positive number <strong>and</strong> define K m byK m =∫ R−R(R 2 − x 2) mdx.Show, using integration by parts, that K m satisfies the recurrence relation(2m +1)K m =2mR 2 K m−1 .(b) For integer n, define I n = K n <strong>and</strong> J n = K n+1/2 . Evaluate I 0 <strong>and</strong> J 0 directly<strong>and</strong> hence prove that(c)I n = 22n+1 (n!) 2 R 2n+1<strong>and</strong> J n =(2n +1)!A sequence of functions V n (R) is defined byV 0 (R) =1,∫ R (√ )V n (R) = V n−1 R2 − x 2−Rπ(2n +1)!R2n+22 2n+1 n!(n +1)! .dx, n ≥ 1.Prove by induction thatV 2n (R) = πn R 2n, V 2n+1 (R) = πn 2 2n+1 n!R 2n+1.n!(2n +1)!(d) For interest,(i) show that V 2n+2 (1)


7Vector algebraThis chapter introduces space vectors <strong>and</strong> their manipulation. Firstly we deal withthe description <strong>and</strong> algebra of vectors, then we consider how vectors may be usedto describe lines <strong>and</strong> planes <strong>and</strong> finally we look at the practical use of vectors infinding distances. Much use of vectors will be made in subsequent chapters; thischapter gives only some basic rules.7.1 Scalars <strong>and</strong> vectorsThe simplest kind of physical quantity is one that can be completely specified byits magnitude, a single number, together with the units in which it is measured.Such a quantity is called a scalar <strong>and</strong> examples include temperature, time <strong>and</strong>density.A vector is a quantity that requires both a magnitude (≥ 0) <strong>and</strong> a direction inspace to specify it completely; we may think of it as an arrow in space. A familiarexample is <strong>for</strong>ce, which has a magnitude (strength) measured in newtons <strong>and</strong> adirection of application. The large number of vectors that are used to describethe physical world include velocity, displacement, momentum <strong>and</strong> electric field.Vectors are also used to describe quantities such as angular momentum <strong>and</strong>surface elements (a surface element has an area <strong>and</strong> a direction defined by thenormal to its tangent plane); in such cases their definitions may seem somewhatarbitrary (though in fact they are st<strong>and</strong>ard) <strong>and</strong> not as physically intuitive as <strong>for</strong>vectors such as <strong>for</strong>ce. A vector is denoted by bold type, the convention of thisbook, or by underlining, the latter being much used in h<strong>and</strong>written work.This chapter considers basic vector algebra <strong>and</strong> illustrates just how powerfulvector analysis can be. All the techniques are presented <strong>for</strong> three-dimensionalspace but most can be readily extended to more dimensions.Throughout the book we will represent a vector in diagrams as a line togetherwith an arrowhead. We will make no distinction between an arrowhead at the212


7.2 ADDITION AND SUBTRACTION OF VECTORSabb + aa + bbaFigure 7.1 Addition of two vectors showing the commutation relation. Wemake no distinction between an arrowhead at the end of the line <strong>and</strong> onealong the line’s length, but rather use that which gives the clearer diagram.end of the line <strong>and</strong> one along the line’s length but, rather, use that which gives theclearer diagram. Furthermore, even though we are considering three-dimensionalvectors, we have to draw them in the plane of the paper. It should not be assumedthat vectors drawn thus are coplanar, unless this is explicitly stated.7.2 Addition <strong>and</strong> subtraction of vectorsThe resultant or vector sum of two displacement vectors is the displacement vectorthat results from per<strong>for</strong>ming first one <strong>and</strong> then the other displacement, as shownin figure 7.1; this process is known as vector addition. However, the principleof addition has physical meaning <strong>for</strong> vector quantities other than displacements;<strong>for</strong> example, if two <strong>for</strong>ces act on the same body then the resultant <strong>for</strong>ce actingon the body is the vector sum of the two. The addition of vectors only makesphysical sense if they are of a like kind, <strong>for</strong> example if they are both <strong>for</strong>cesacting in three dimensions. It may be seen from figure 7.1 that vector addition iscommutative, i.e.a + b = b + a. (7.1)The generalisation of this procedure to the addition of three (or more) vectors isclear <strong>and</strong> leads to the associativity property of addition (see figure 7.2), e.g.a +(b + c) =(a + b)+c. (7.2)Thus, it is immaterial in what order any number of vectors are added.The subtraction of two vectors is very similar to their addition (see figure 7.3),that is,a − b = a +(−b)where −b is a vector of equal magnitude but exactly opposite direction to vector b.213


VECTOR ALGEBRAabcb + cbcab + ca +(b + c)aba + ba + b c(a + b)+cFigure 7.2Addition of three vectors showing the associativity relation.a−ba − babFigure 7.3Subtraction of two vectors.The subtraction of two equal vectors yields the zero vector, 0, which has zeromagnitude <strong>and</strong> no associated direction.7.3 Multiplication by a scalarMultiplication of a vector by a scalar (not to be confused with the ‘scalarproduct’, to be discussed in subsection 7.6.1) gives a vector in the same directionas the original but of a proportional magnitude. This can be seen in figure 7.4.The scalar may be positive, negative or zero. It can also be complex in someapplications. Clearly, when the scalar is negative we obtain a vector pointingin the opposite direction to the original vector. Multiplication by a scalar isassociative, commutative <strong>and</strong> distributive over addition. These properties may besummarised <strong>for</strong> arbitrary vectors a <strong>and</strong> b <strong>and</strong> arbitrary scalars λ <strong>and</strong> µ by(λµ)a = λ(µa) =µ(λa), (7.3)λ(a + b) =λa + λb, (7.4)(λ + µ)a = λa + µa. (7.5)214


7.3 MULTIPLICATION BY A SCALARλ aaFigure 7.4Scalar multiplication of a vector (<strong>for</strong> λ>1).BµbpPλAOaFigure 7.5 An illustration of the ratio theorem. The point P divides the linesegment AB in the ratio λ : µ.Having defined the operations of addition, subtraction <strong>and</strong> multiplication by ascalar, we can now use vectors to solve simple problems in geometry.◮A point P divides a line segment AB in the ratio λ : µ (see figure 7.5). If the positionvectors of the points A <strong>and</strong> B are a <strong>and</strong> b, respectively, find the position vector of thepoint P .As is conventional <strong>for</strong> vector geometry problems, we denote the vector from the point Ato the point B by AB. If the position vectors of the points A <strong>and</strong> B, relative to some originO, area <strong>and</strong> b, it should be clear that AB = b − a.Now, from figure 7.5 we see that one possible way of reaching the point P from O isfirst to go from O to A <strong>and</strong> to go along the line AB <strong>for</strong> a distance equal to the the fractionλ/(λ + µ) of its total length. We may express this in terms of vectors asOP = p = a +λλ + µ AB= a + λ (b − a)=λ + µ(1 − λλ + µ= µλ + µ a +)a +λλ + µ bλ b, (7.6)λ + µwhich expresses the position vector of the point P in terms of those of A <strong>and</strong> B. Wewould,of course, obtain the same result by considering the path from O to B <strong>and</strong> then to P . ◭215


VECTOR ALGEBRACEAGFDacBbOFigure 7.6 The centroid of a triangle. The triangle is defined by the points A,B <strong>and</strong> C that have position vectors a, b <strong>and</strong> c. The broken lines CD, BE, AFconnect the vertices of the triangle to the mid-points of the opposite sides;these lines intersect at the centroid G of the triangle.Result (7.6) is a version of the ratio theorem <strong>and</strong> we may use it in solving morecomplicated problems.◮The vertices of triangle ABC have position vectors a, b <strong>and</strong> c relative to some origin O(see figure 7.6). Find the position vector of the centroid G of the triangle.From figure 7.6, the points D <strong>and</strong> E bisect the lines AB <strong>and</strong> AC respectively. Thus fromthe ratio theorem (7.6), with λ = µ =1/2, the position vectors of D <strong>and</strong> E relative to theorigin ared = 1 a + 1 b,2 2e = 1 a + 1 c.2 2Using the ratio theorem again, we may write the position vector of a general point on theline CD that divides the line in the ratio λ :(1− λ) asr =(1− λ)c + λd,=(1− λ)c + 1 λ(a + b), (7.7)2where we have expressed d in terms of a <strong>and</strong> b. Similarly, the position vector of a generalpoint on the line BE can be expressed asr =(1− µ)b + µe,=(1− µ)b + 1 µ(a + c). (7.8)2Thus, at the intersection of the lines CD <strong>and</strong> BE we require, from (7.7), (7.8),(1 − λ)c + 1 λ(a + b) =(1− µ)b + 1 µ(a + c).2 2By equating the coefficents of the vectors a, b, c we find1λ = µ, λ =1− µ, 1 − λ = 1 µ.2 2216


7.4 BASIS VECTORS AND COMPONENTSThese equations are consistent <strong>and</strong> have the solution λ = µ =2/3. Substituting thesevalues into either (7.7) or (7.8) we find that the position vector of the centroid G is givenbyg = 1 (a + b + c). ◭37.4 Basis vectors <strong>and</strong> componentsGiven any three different vectors e 1 , e 2 <strong>and</strong> e 3 , which do not all lie in a plane,it is possible, in a three-dimensional space, to write any other vector in terms ofscalar multiples of them:a = a 1 e 1 + a 2 e 2 + a 3 e 3 . (7.9)The three vectors e 1 , e 2 <strong>and</strong> e 3 are said to <strong>for</strong>m a basis (<strong>for</strong> the three-dimensionalspace); the scalars a 1 , a 2 <strong>and</strong> a 3 , which may be positive, negative or zero, arecalled the components of the vector a with respect to this basis. We say that thevector has been resolved into components.Most often we shall use basis vectors that are mutually perpendicular, <strong>for</strong> easeof manipulation, though this is not necessary. In general, a basis set must(i) have as many basis vectors as the number of dimensions (in more <strong>for</strong>mallanguage, the basis vectors must span the space) <strong>and</strong>(ii) be such that no basis vector may be described as a sum of the others, or,more <strong>for</strong>mally, the basis vectors must be linearly independent. Putting thismathematically, in N dimensions, we requirec 1 e 1 + c 2 e 2 + ···+ c N e N ≠ 0,<strong>for</strong> any set of coefficients c 1 ,c 2 ,...,c N except c 1 = c 2 = ···= c N =0.In this chapter we will only consider vectors in three dimensions; higher dimensionalitycan be achieved by simple extension.If we wish to label points in space using a Cartesian coordinate system (x, y, z),we may introduce the unit vectors i, j <strong>and</strong> k, which point along the positive x-,y- <strong>and</strong>z- axes respectively. A vector a may then be written as a sum of threevectors, each parallel to a different coordinate axis:a = a x i + a y j + a z k. (7.10)A vector in three-dimensional space thus requires three components to describefully both its direction <strong>and</strong> its magnitude. A displacement in space may bethought of as the sum of displacements along the x-, y- <strong>and</strong>z- directions (seefigure 7.7). For brevity, the components of a vector a with respect to a particularcoordinate system are sometimes written in the <strong>for</strong>m (a x ,a y ,a z ). Note that the217


VECTOR ALGEBRAkaja z ka y ja x iiFigure 7.7 A Cartesian basis set. The vector a is the sum of a x i, a y j <strong>and</strong> a z k.basis vectors i, j <strong>and</strong> k may themselves be represented by (1, 0, 0), (0, 1, 0) <strong>and</strong>(0, 0, 1) respectively.We can consider the addition <strong>and</strong> subtraction of vectors in terms of theircomponents. The sum of two vectors a <strong>and</strong> b is found by simply adding theircomponents, i.e.a + b = a x i + a y j + a z k + b x i + b y j + b z k=(a x + b x )i +(a y + b y )j +(a z + b z )k, (7.11)<strong>and</strong> their difference by subtracting them,a − b = a x i + a y j + a z k − (b x i + b y j + b z k)=(a x − b x )i +(a y − b y )j +(a z − b z )k. (7.12)◮Two particles have velocities v 1 = i +3j +6k <strong>and</strong> v 2 = i − 2k, respectively. Find thevelocity u of the second particle relative to the first.The required relative velocity is given byu = v 2 − v 1 =(1− 1)i +(0− 3)j +(−2 − 6)k= −3j − 8k. ◭7.5 Magnitude of a vectorThe magnitude of the vector a is denoted by |a| or a. In terms of its componentsin three-dimensional Cartesian coordinates, the magnitude of a is given by√a ≡|a| = a 2 x + a 2 y + a 2 z. (7.13)Hence, the magnitude of a vector is a measure of its length. Such an analogy isuseful <strong>for</strong> displacement vectors but magnitude is better described, <strong>for</strong> example, by‘strength’ <strong>for</strong> vectors such as <strong>for</strong>ce or by ‘speed’ <strong>for</strong> velocity vectors. For instance,218


7.6 MULTIPLICATION OF VECTORSbOθ b cos θaFigure 7.8 The projection of b onto the direction of a is b cos θ. The scalarproduct of a <strong>and</strong> b is ab cos θ.in the previous example, the speed of the second particle relative to the first isgiven byu = |u| = √ (−3) 2 +(−8) 2 = √ 73.A vector whose magnitude equals unity is called a unit vector. The unit vectorin the direction a is usually notated â <strong>and</strong> may be evaluated asâ = a|a| . (7.14)The unit vector is a useful concept because a vector written as λâ then has magnitudeλ <strong>and</strong> direction â. Thus magnitude <strong>and</strong> direction are explicitly separated.7.6 Multiplication of vectorsWe have already considered multiplying a vector by a scalar. Now we considerthe concept of multiplying one vector by another vector. It is not immediatelyobvious what the product of two vectors represents <strong>and</strong> in fact two productsare commonly defined, the scalar product <strong>and</strong> the vector product. As their namesimply, the scalar product of two vectors is just a number, whereas the vectorproduct is itself a vector. Although neither the scalar nor the vector productis what we might normally think of as a product, their use is widespread <strong>and</strong>numerous examples will be described elsewhere in this book.7.6.1 Scalar productThe scalar product (or dot product) of two vectors a <strong>and</strong> b is denoted by a · b<strong>and</strong> is given bya · b ≡|a||b| cos θ, 0 ≤ θ ≤ π, (7.15)where θ is the angle between the two vectors, placed ‘tail to tail’ or ‘head to head’.Thus, the value of the scalar product a · b equals the magnitude of a multipliedby the projection of b onto a (see figure 7.8).219


VECTOR ALGEBRAFrom (7.15) we see that the scalar product has the particularly useful propertythata · b = 0 (7.16)is a necessary <strong>and</strong> sufficient condition <strong>for</strong> a to be perpendicular to b (unless eitherof them is zero). It should be noted in particular that the Cartesian basis vectorsi, j <strong>and</strong> k, being mutually orthogonal unit vectors, satisfy the equationsi · i = j · j = k · k =1, (7.17)i · j = j · k = k · i =0. (7.18)Examples of scalar products arise naturally throughout physics <strong>and</strong> in particularin connection with energy. Perhaps the simplest is the work done F · r inmoving the point of application of a constant <strong>for</strong>ce F through a displacement r;notice that, as expected, if the displacement is perpendicular to the direction ofthe <strong>for</strong>ce then F · r = 0 <strong>and</strong> no work is done. A second simple example is af<strong>for</strong>dedby the potential energy −m · B of a magnetic dipole, represented in strength <strong>and</strong>orientation by a vector m, placed in an external magnetic field B.As the name implies, the scalar product has a magnitude but no direction. Thescalar product is commutative <strong>and</strong> distributive over addition:a · b = b · a (7.19)a · (b + c) =a · b + a · c. (7.20)◮Four non-coplanar points A, B, C, D are positioned such that the line AD is perpendicularto BC <strong>and</strong> BD is perpendicular to AC. Show that CD is perpendicular to AB.Denote the four position vectors by a, b, c, d. As none of the three pairs of lines actuallyintersect, it is difficult to indicate their orthogonality in the diagram we would normallydraw. However, the orthogonality can be expressed in vector <strong>for</strong>m <strong>and</strong> we start by notingthat, since AD ⊥ BC, it follows from (7.16) that(d − a) · (c − b) =0.Similarly, since BD ⊥ AC,(d − b) · (c − a) =0.Combining these two equations we find(d − a) · (c − b) =(d − b) · (c − a),which, on mutliplying out the parentheses, givesd · c − a · c − d · b + a · b = d · c − b · c − d · a + b · a.Cancelling terms that appear on both sides <strong>and</strong> rearranging yieldsd · b − d · a − c · b + c · a =0,which simplifies to give(d − c) · (b − a) =0.From (7.16), we see that this implies that CD is perpendicular to AB. ◭220


7.6 MULTIPLICATION OF VECTORSIf we introduce a set of basis vectors that are mutually orthogonal, such as i, j,k, we can write the components of a vector a, with respect to that basis, in termsof the scalar product of a with each of the basis vectors, i.e. a x = a·i, a y = a·j <strong>and</strong>a z = a · k. In terms of the components a x , a y <strong>and</strong> a z the scalar product is given bya · b =(a x i + a y j + a z k) · (b x i + b y j + b z k)=a x b x + a y b y + a z b z , (7.21)where the cross terms such as a x i · b y j are zero because the basis vectors aremutually perpendicular; see equation (7.18). It should be clear from (7.15) thatthe value of a · b has a geometrical definition <strong>and</strong> that this value is independentof the actual basis vectors used.◮Find the angle between the vectors a = i +2j +3k <strong>and</strong> b =2i +3j +4k.From (7.15) the cosine of the angle θ between a <strong>and</strong> b is given bycos θ = a · b|a||b| .From (7.21) the scalar product a · b has the valuea · b =1× 2+2× 3+3× 4=20,<strong>and</strong> from (7.13) the lengths of the vectors areThus,|a| = √ 1 2 +2 2 +3 2 = √ 14 <strong>and</strong> |b| = √ 2 2 +3 2 +4 2 = √ 29.cos θ =20√14√29≈ 0.9926 ⇒ θ =0.12 rad. ◭We can see from the expressions (7.15) <strong>and</strong> (7.21) <strong>for</strong> the scalar product that ifθ is the angle between a <strong>and</strong> b thencos θ = a x b xa b + a y b ya b + a z b za bwhere a x /a, a y /a <strong>and</strong> a z /a are called the direction cosines of a, since they give thecosine of the angle made by a with each of the basis vectors. Similarly b x /b, b y /b<strong>and</strong> b z /b are the direction cosines of b.If we take the scalar product of any vector a with itself then clearly θ = 0 <strong>and</strong>from (7.15) we havea · a = |a| 2 .Thus the magnitude of a can be written in a coordinate-independent <strong>for</strong>m as|a| = √ a · a.Finally, we note that the scalar product may be extended to vectors withcomplex components if it is redefined asa · b = a ∗ xb x + a ∗ yb y + a ∗ zb z ,where the asterisk represents the operation of complex conjugation. To accom-221


VECTOR ALGEBRAa × bθbaFigure 7.9set.The vector product. The vectors a, b <strong>and</strong> a×b <strong>for</strong>m a right-h<strong>and</strong>edmodate this extension the commutation property (7.19) must be modified toreada · b =(b · a) ∗ . (7.22)In particular it should be noted that (λa) · b = λ ∗ a · b, whereas a · (λb) =λa · b.However, the magnitude of a complex vector is still given by |a| = √ a · a, sincea · a is always real.7.6.2 Vector productThe vector product (or cross product) of two vectors a <strong>and</strong> b is denoted by a × b<strong>and</strong> is defined to be a vector of magnitude |a||b| sin θ in a direction perpendicularto both a <strong>and</strong> b;|a × b| = |a||b| sin θ.The direction is found by ‘rotating’ a into b through the smallest possible angle.The sense of rotation is that of a right-h<strong>and</strong>ed screw that moves <strong>for</strong>ward in thedirection a × b (see figure 7.9). Again, θ is the angle between the two vectorsplaced ‘tail to tail’ or ‘head to head’. With this definition a, b <strong>and</strong> a × b <strong>for</strong>m aright-h<strong>and</strong>ed set. A more directly usable description of the relative directions ina vector product is provided by a right h<strong>and</strong> whose first two fingers <strong>and</strong> thumbare held to be as nearly mutually perpendicular as possible. If the first finger ispointed in the direction of the first vector <strong>and</strong> the second finger in the directionof the second vector, then the thumb gives the direction of the vector product.The vector product is distributive over addition, but anticommutative <strong>and</strong> nonassociative:(a + b) × c =(a × c)+(b × c), (7.23)b × a = −(a × b), (7.24)(a × b) × c ≠ a × (b × c). (7.25)222


7.6 MULTIPLICATION OF VECTORSPFθRrOFigure 7.10 The moment of the <strong>for</strong>ce F about O is r×F. The cross representsthe direction of r × F, which is perpendicularly into the plane of the paper.From its definition, we see that the vector product has the very useful propertythat if a × b = 0 then a is parallel or antiparallel to b (unless either of them iszero). We also note thata × a = 0. (7.26)◮Show that if a = b + λc, <strong>for</strong> some scalar λ, thena × c = b × c.From (7.23) we havea × c =(b + λc) × c = b × c + λc × c.However, from (7.26), c × c = 0 <strong>and</strong> soa × c = b × c. (7.27)We note in passing that the fact that (7.27) is satisfied does not imply that a = b. ◭An example of the use of the vector product is that of finding the area, A, ofa parallelogram with sides a <strong>and</strong> b, using the <strong>for</strong>mulaA = |a × b|. (7.28)Another example is af<strong>for</strong>ded by considering a <strong>for</strong>ce F acting through a point R,whose vector position relative to the origin O is r (see figure 7.10). Its momentor torque about O is the strength of the <strong>for</strong>ce times the perpendicular distanceOP, which numerically is just Fr sin θ, i.e. the magnitude of r × F. Furthermore,the sense of the moment is clockwise about an axis through O that pointsperpendicularly into the plane of the paper (the axis is represented by a crossin the figure). Thus the moment is completely represented by the vector r × F,in both magnitude <strong>and</strong> spatial sense. It should be noted that the same vectorproduct is obtained wherever the point R is chosen, so long as it lies on the lineof action of F.Similarly, if a solid body is rotating about some axis that passes through theorigin, with an angular velocity ω then we can describe this rotation by a vectorω that has magnitude ω <strong>and</strong> points along the axis of rotation. The direction of ω223


VECTOR ALGEBRAis the <strong>for</strong>ward direction of a right-h<strong>and</strong>ed screw rotating in the same sense as thebody. The velocity of any point in the body with position vector r is then givenby v = ω × r.Since the basis vectors i, j, k are mutually perpendicular unit vectors, <strong>for</strong>minga right-h<strong>and</strong>ed set, their vector products are easily seen to bei × i = j × j = k × k = 0, (7.29)i × j = −j × i = k, (7.30)j × k = −k × j = i, (7.31)k × i = −i × k = j. (7.32)Using these relations, it is straight<strong>for</strong>ward to show that the vector product of twogeneral vectors a <strong>and</strong> b is given in terms of their components with respect to thebasis set i, j, k, bya × b =(a y b z − a z b y )i +(a z b x − a x b z )j +(a x b y − a y b x )k. (7.33)For the reader who is familiar with determinants (see chapter 8), we record thatthis can also be written as∣ i j k ∣∣∣∣∣a × b =a x a y a z .∣ b x b y b zThat the cross product a × b is perpendicular to both a <strong>and</strong> b can be verifiedin component <strong>for</strong>m by <strong>for</strong>ming its dot products with each of the two vectors <strong>and</strong>showing that it is zero in both cases.◮Find the area A of the parallelogram with sides a = i +2j +3k <strong>and</strong> b =4i +5j +6k.The vector product a × b is given in component <strong>for</strong>m bya × b =(2× 6 − 3 × 5)i +(3× 4 − 1 × 6)j +(1× 5 − 2 × 4)k= −3i +6j − 3k.Thus the area of the parallelogram isA = |a × b| = √ (−3) 2 +6 2 +(−3) 2 = √ 54. ◭7.6.3 Scalar triple productNow that we have defined the scalar <strong>and</strong> vector products, we can extend ourdiscussion to define products of three vectors. Again, there are two possibilities,the scalar triple product <strong>and</strong> the vector triple product.224


7.6 MULTIPLICATION OF VECTORSvPOφθcbaFigure 7.11The scalar triple product gives the volume of a parallelepiped.The scalar triple product is denoted by[a, b, c] ≡ a · (b × c)<strong>and</strong>, as its name suggests, it is just a number. It is most simply interpreted as thevolume of a parallelepiped whose edges are given by a, b <strong>and</strong> c (see figure 7.11).The vector v = a × b is perpendicular to the base of the solid <strong>and</strong> has magnitudev = ab sin θ, i.e. the area of the base. Further, v · c = vc cos φ. Thus, since c cos φ= OP is the vertical height of the parallelepiped, it is clear that (a × b) · c =areaof the base × perpendicular height = volume. It follows that, if the vectors a, b<strong>and</strong> c are coplanar, a · (b × c) =0.Expressed in terms of the components of each vector with respect to theCartesian basis set i, j, k the scalar triple product isa · (b × c) =a x (b y c z − b z c y )+a y (b z c x − b x c z )+a z (b x c y − b y c x ),(7.34)which can also be written as a determinant:∣ a x a y a z ∣∣∣∣∣a · (b × c) =b x b y b z .∣ c x c y c zBy writing the vectors in component <strong>for</strong>m, it can be shown thata · (b × c) =(a × b) · c,so that the dot <strong>and</strong> cross symbols can be interchanged without changing the result.More generally, the scalar triple product is unchanged under cyclic permutationof the vectors a, b, c. Other permutations simply give the negative of the originalscalar triple product. These results can be summarised by[a, b, c] =[b, c, a] =[c, a, b] =−[a, c, b] =−[b, a, c] =−[c, b, a]. (7.35)225


VECTOR ALGEBRA◮Find the volume V of the parallelepiped with sides a = i +2j +3k, b =4i +5j +6k <strong>and</strong>c =7i +8j +10k.We have already found that a × b = −3i +6j − 3k, in subsection 7.6.2. Hence the volumeof the parallelepiped is given byV = |a · (b × c)| = |(a × b) · c|= |(−3i +6j − 3k) · (7i +8j +10k)|= |(−3)(7) + (6)(8) + (−3)(10)| =3. ◭Another useful <strong>for</strong>mula involving both the scalar <strong>and</strong> vector products is Lagrange’sidentity (see exercise 7.9), i.e.(a × b) · (c × d) ≡ (a · c)(b · d) − (a · d)(b · c). (7.36)7.6.4 Vector triple productBy the vector triple product of three vectors a, b, c we mean the vector a × (b × c).Clearly, a × (b × c) is perpendicular to a <strong>and</strong> lies in the plane of b <strong>and</strong> c <strong>and</strong> socan be expressed in terms of them (see (7.37) below). We note, from (7.25), thatthe vector triple product is not associative, i.e. a × (b × c) ≠(a × b) × c.Two useful <strong>for</strong>mulae involving the vector triple product area × (b × c) =(a · c)b − (a · b)c, (7.37)(a × b) × c =(a · c)b − (b · c)a, (7.38)which may be derived by writing each vector in component <strong>for</strong>m (see exercise 7.8).It can also be shown that <strong>for</strong> any three vectors a, b, c,a × (b × c)+b × (c × a)+c × (a × b) =0.7.7 Equations of lines, planes <strong>and</strong> spheresNow that we have described the basic algebra of vectors, we can apply the resultsto a variety of problems, the first of which is to find the equation of a line invector <strong>for</strong>m.7.7.1 Equation of a lineConsider the line passing through the fixed point A with position vector a <strong>and</strong>having a direction b (see figure 7.12). It is clear that the position vector r of ageneral point R on the line can be written asr = a + λb, (7.39)226


7.7 EQUATIONS OF LINES, PLANES AND SPHERESRAbraOFigure 7.12 The equation of a line. The vector b is in the direction AR <strong>and</strong>λb is the vector from A to R.since R can be reached by starting from O, going along the translation vectora to the point A on the line <strong>and</strong> then adding some multiple λb of the vector b.Different values of λ give different points R on the line.Taking the components of (7.39), we see that the equation of the line can alsobewritteninthe<strong>for</strong>mx − a xb x= y − a yb y= z − a zb z= constant. (7.40)Taking the vector product of (7.39) with b <strong>and</strong> remembering that b × b = 0 givesan alternative equation <strong>for</strong> the line(r − a) × b = 0.We may also find the equation of the line that passes through two fixed pointsA <strong>and</strong> C with position vectors a <strong>and</strong> c. SinceAC is given by c − a, the positionvector of a general point on the line isr = a + λ(c − a).7.7.2 Equation of a planeThe equation of a plane through a point A with position vector a <strong>and</strong> perpendicularto a unit position vector ˆn (see figure 7.13) is(r − a) · ˆn =0. (7.41)This follows since the vector joining A to a general point R with position vectorr is r − a; r will lie in the plane if this vector is perpendicular to the normal tothe plane. Rewriting (7.41) as r · ˆn = a · ˆn, we see that the equation of the planemay also be expressed in the <strong>for</strong>m r · ˆn = d, or in component <strong>for</strong>m aslx + my + nz = d, (7.42)227


VECTOR ALGEBRAˆnARadrOFigure 7.13 The equation of the plane is (r − a) · ˆn =0.where the unit normal to the plane is ˆn = li + mj + nk <strong>and</strong> d = a · ˆn is theperpendicular distance of the plane from the origin.The equation of a plane containing points a, b <strong>and</strong> c isr = a + λ(b − a)+µ(c − a).This is apparent because starting from the point a in the plane, all other pointsmay be reached by moving a distance along each of two (non-parallel) directionsin the plane. Two such directions are given by b − a <strong>and</strong> c − a. It can be shownthat the equation of this plane may also be written in the more symmetrical <strong>for</strong>mwhere α + β + γ =1.r = αa + βb + γc,◮Find the direction of the line of intersection of the two planes x +3y − z = 5 <strong>and</strong>2x − 2y +4z =3.The two planes have normal vectors n 1 = i +3j − k <strong>and</strong> n 2 =2i − 2j +4k. Itisclearthat these are not parallel vectors <strong>and</strong> so the planes must intersect along some line. Thedirection p of this line must be parallel to both planes <strong>and</strong> hence perpendicular to bothnormals. There<strong>for</strong>ep = n 1 × n 2= [(3)(4) − (−2)(−1)] i +[(−1)(2) − (1)(4)] j + [(1)(−2) − (3)(2)] k=10i − 6j − 8k. ◭7.7.3 Equation of a sphereClearly, the defining property of a sphere is that all points on it are equidistantfrom a fixed point in space <strong>and</strong> that the common distance is equal to the radius228


7.8 USING VECTORS TO FIND DISTANCESof the sphere. This is easily expressed in vector notation as|r − c| 2 =(r − c) · (r − c) =a 2 , (7.43)where c is the position vector of the centre of the sphere <strong>and</strong> a is its radius.◮Find the radius ρ of the circle that is the intersection of the plane ˆn · r = p <strong>and</strong> the sphereof radius a centred on the point with position vector c.The equation of the sphere is|r − c| 2 = a 2 , (7.44)<strong>and</strong> that of the circle of intersection is|r − b| 2 = ρ 2 , (7.45)where r is restricted to lie in the plane <strong>and</strong> b is the position of the circle’s centre.As b lies on the plane whose normal is ˆn, thevectorb − c must be parallel to ˆn, i.e.b − c = λˆn <strong>for</strong> some λ. Further, by Pythagoras, we must have ρ 2 + |b − c| 2 = a 2 . Thusλ 2 = a 2 − ρ 2 .Writing b = c + √ a 2 − ρ 2 ˆn <strong>and</strong> substituting in (7.45) gives(r 2 − 2r · c + √ )a 2 − ρ 2 ˆn + c 2 +2(c · ˆn) √ a 2 − ρ 2 + a 2 − ρ 2 = ρ 2 ,whilst, on expansion, (7.44) becomesr 2 − 2r · c + c 2 = a 2 .Subtracting these last two equations, using ˆn · r = p <strong>and</strong> simplifying yieldsp − c · ˆn = √ a 2 − ρ 2 .On rearrangement, this gives ρ as √ a 2 − (p − c · ˆn) 2 , which places obvious geometricalconstraints on the values a, c, ˆn <strong>and</strong> p can take if a real intersection between the sphere<strong>and</strong> the plane is to occur. ◭7.8 Using vectors to find distancesThis section deals with the practical application of vectors to finding distances.Some of these problems are extremely cumbersome in component <strong>for</strong>m, but theyall reduce to neat solutions when general vectors, with no explicit basis set,are used. These examples show the power of vectors in simplifying geometricalproblems.7.8.1 Distance from a point to a lineFigure 7.14 shows a line having direction b that passes through a point A whoseposition vector is a. To find the minimum distance d of the line from a point Pwhose position vector is p, we must solve the right-angled triangle shown. We seethat d = |p − a| sin θ; so, from the definition of the vector product, it follows thatd = |(p − a) × ˆb|.229


VECTOR ALGEBRAPp − apdAbθaOFigure 7.14The minimum distance from a point to a line.◮Find the minimum distance from the point P with coordinates (1, 2, 1) to the line r = a+λb,where a = i + j + k <strong>and</strong> b =2i − j +3k.Comparison with (7.39) shows that the line passes through the point (1, 1, 1) <strong>and</strong> hasdirection 2i − j +3k. The unit vector in this direction isˆb = √ 1 (2i − j +3k).14The position vector of P is p = i +2j + k <strong>and</strong> we find(p − a) × ˆb = √ 1 [ j × (2i − 3j +3k)]14= √ 1 (3i − 2k).14Thus the minimum distance from the line to the point P is d = √ 13/14. ◭7.8.2 Distance from a point to a planeThe minimum distance d from a point P whose position vector is p to the planedefined by (r − a) · ˆn = 0 may be deduced by finding any vector from P to theplane <strong>and</strong> then determining its component in the normal direction. This is shownin figure 7.15. Consider the vector a − p, which is a particular vector from P tothe plane. Its component normal to the plane, <strong>and</strong> hence its distance from theplane, is given byd =(a − p) · ˆn, (7.46)where the sign of d depends on which side of the plane P is situated.230


7.8 USING VECTORS TO FIND DISTANCESPpˆndaOFigure 7.15The minimum distance d from a point to a plane.◮Find the distance from the point P with coordinates (1, 2, 3) to the plane that contains thepoints A, B <strong>and</strong> C having coordinates (0, 1, 0), (2, 3, 1) <strong>and</strong> (5, 7, 2).Let us denote the position vectors of the points A, B, C by a, b, c. Two vectors in theplane areb − a =2i +2j + k <strong>and</strong> c − a =5i +6j +2k,<strong>and</strong> hence a vector normal to the plane isn =(2i +2j + k) × (5i +6j +2k) =−2i + j +2k,<strong>and</strong> its unit normal isˆn = n|n| = 1 (−2i + j +2k).3Denoting the position vector of P by p, the minimum distance from the plane to P isgiven byd =(a − p) · ˆn=(−i − j − 3k) · 1 (−2i + j +2k)3= 2 − 1 − 2 = − 5 .3 3 3If we take P to be the origin O, then we find d = 1 , i.e. a positive quantity. It follows from3this that the original point P with coordinates (1, 2, 3), <strong>for</strong> which d was negative, is on theopposite side of the plane from the origin. ◭7.8.3 Distance from a line to a lineConsider two lines in the directions a <strong>and</strong> b, as shown in figure 7.16. Since a × bis by definition perpendicular to both a <strong>and</strong> b, the unit vector normal to boththese lines isˆn = a × b|a × b| .231


VECTOR ALGEBRAQbqˆnPpaOFigure 7.16The minimum distance from one line to another.If p <strong>and</strong> q are the position vectors of any two points P <strong>and</strong> Q on different linesthen the vector connecting them is p − q. Thus, the minimum distance d betweenthe lines is this vector’s component along the unit normal, i.e.d = |(p − q) · ˆn|.◮A line is inclined at equal angles to the x-, y- <strong>and</strong> z-axes <strong>and</strong> passes through the origin.Another line passes through the points (1, 2, 4) <strong>and</strong> (0, 0, 1). Find the minimum distancebetween the two lines.The first line is given byr 1 = λ(i + j + k),<strong>and</strong> the second byr 2 = k + µ(i +2j +3k).Hence a vector normal to both lines isn =(i + j + k) × (i +2j +3k) =i − 2j + k,<strong>and</strong> the unit normal isˆn = √ 1 (i − 2j + k).6A vector between the two lines is, <strong>for</strong> example, the one connecting the points (0, 0, 0)<strong>and</strong> (0, 0, 1), which is simply k. Thus it follows that the minimum distance between thetwo lines isd = √ 1 |k · (i − 2j + k)| = √ 1 . ◭6 67.8.4 Distance from a line to a planeLet us consider the line r = a + λb. This line will intersect any plane to which itis not parallel. Thus, if a plane has a normal ˆn then the minimum distance from232


7.9 RECIPROCAL VECTORSthe line to the plane is zero unlessin which case the distance, d, will bewhere r is any point in the plane.b · ˆn =0,d = |(a − r) · ˆn|,◮A line is given by r = a + λb, wherea = i +2j +3k <strong>and</strong> b =4i +5j +6k. Findthecoordinates of the point P at which the line intersects the planex +2y +3z =6.A vector normal to the plane isn = i +2j +3k,from which we find that b · n ≠ 0. Thus the line does indeed intersect the plane. To findthe point of intersection we merely substitute the x-, y- <strong>and</strong>z- values of a general pointon the line into the equation of the plane, obtaining1+4λ +2(2+5λ)+3(3+6λ) =6 ⇒ 14 + 32λ =6.This gives λ = − 1 , which we may substitute into the equation <strong>for</strong> the line to obtain4x =1− 1 (4) = 0, y =2− 1 (5) = 3 <strong>and</strong> z =3− 1 (6) = 3 . Thus the point of intersection is4 4 4 4 2(0, 3 4 , 3 2 ). ◭ 7.9 Reciprocal vectorsThe final section of this chapter introduces the concept of reciprocal vectors,which have particular uses in crystallography.The two sets of vectors a, b, c <strong>and</strong> a ′ , b ′ , c ′ are called reciprocal sets if<strong>and</strong>a · a ′ = b · b ′ = c · c ′ = 1 (7.47)a ′ · b = a ′ · c = b ′ · a = b ′ · c = c ′ · a = c ′ · b =0. (7.48)It can be verified (see exercise 7.19) that the reciprocal vectors of a, b <strong>and</strong> c aregiven bya ′ =b × ca · (b × c) , (7.49)b ′ =c × aa · (b × c) , (7.50)c ′ =a × ba · (b × c) , (7.51)where a · (b × c) ≠ 0. In other words, reciprocal vectors only exist if a, b <strong>and</strong> c are233


VECTOR ALGEBRAnot coplanar. Moreover, if a, b <strong>and</strong> c are mutually orthogonal unit vectors thena ′ = a, b ′ = b <strong>and</strong> c ′ = c, so that the two systems of vectors are identical.◮Construct the reciprocal vectors of a =2i, b = j + k, c = i + k.First we evaluate the triple scalar product:a · (b × c) =2i · [(j + k) × (i + k)]=2i · (i + j − k) = 2.Now we find the reciprocal vectors:a ′ = 1 (j + k) × (i + k) = 1 (i + j − k),2 2b ′ = 1 (i + k) × 2i = j,2c ′ = 1 (2i) × (j + k) = −j + k.2It is easily verified that these reciprocal vectors satisfy their defining properties (7.47),(7.48). ◭We may also use the concept of reciprocal vectors to define the components of avector a with respect to basis vectors e 1 , e 2 , e 3 that are not mutually orthogonal.If the basis vectors are of unit length <strong>and</strong> mutually orthogonal, such as theCartesian basis vectors i, j, k, then (see the text preceeding (7.21)) the vector acanbewritteninthe<strong>for</strong>ma =(a · i)i +(a · j)j +(a · k)k.If the basis is not orthonormal, however, then this is no longer true. Nevertheless,we may write the components of a with respect to a non-orthonormal basise 1 , e 2 , e 3 in terms of its reciprocal basis vectors e ′ 1 , e′ 2 , e′ 3 , which are defined as in(7.49)–(7.51). If we leta = a 1 e 1 + a 2 e 2 + a 3 e 3 ,then the scalar product a · e ′ 1 is given bya · e ′ 1 = a 1 e 1 · e ′ 1 + a 2 e 2 · e ′ 1 + a 3 e 3 · e ′ 1 = a 1 ,where we have used the relations (7.48). Similarly, a 2 = a·e ′ 2 <strong>and</strong> a 3 = a·e ′ 3 ;sonowa =(a · e ′ 1)e 1 +(a · e ′ 2)e 2 +(a · e ′ 3)e 3 . (7.52)7.10 Exercises7.1 Which of the following statements about general vectors a, b <strong>and</strong> c are true?(a) c · (a × b) =(b × a) · c.(b) a × (b × c) =(a × b) × c.(c) a × (b × c) =(a · c)b − (a · b)c.(d) d = λa + µb implies (a × b) · d =0.(e) a × c = b × c implies c · a − c · b = c|a − b|.(f) (a × b) × (c × b) =b[b · (c × a)].234


7.10 EXERCISES7.2 A unit cell of diamond is a cube of side A, with carbon atoms at each corner, atthe centre of each face <strong>and</strong>, in addition, at positions displaced by 1 A(i + j + k)4from each of those already mentioned; i, j, k are unit vectors along the cube axes.One corner of the cube is taken as the origin of coordinates. What are the vectorsjoining the atom at 1 A(i + j + k) to its four nearest neighbours? Determine the4angle between the carbon bonds in diamond.7.3 Identify the following surfaces:(a) |r| = k; (b)r · u = l; (c)r · u = m|r| <strong>for</strong> −1 ≤ m ≤ +1;(d) |r − (r · u)u| = n.Here k, l, m <strong>and</strong> n are fixed scalars <strong>and</strong> u is a fixed unit vector.7.4 Find the angle between the position vectors to the points (3, −4, 0) <strong>and</strong> (−2, 1, 0)<strong>and</strong> find the direction cosines of a vector perpendicular to both.7.5 A, B, C <strong>and</strong> D are the four corners, in order, of one face of a cube of side 2units. The opposite face has corners E,F,G <strong>and</strong> H, withAE, BF, CG <strong>and</strong> DHas parallel edges of the cube. The centre O ofthecubeistakenastheorigin<strong>and</strong> the x-, y- <strong>and</strong>z-axes are parallel to AD, AE <strong>and</strong> AB, respectively. Find thefollowing:(a) the angle between the face diagonal AF <strong>and</strong> the body diagonal AG;(b) the equation of the plane through B that is parallel to the plane CGE;(c) the perpendicular distance from the centre J of the face BCGF to the planeOCG;(d) the volume of the tetrahedron JOCG.7.6 Use vector methods to prove that the lines joining the mid-points of the oppositeedges of a tetrahedron OABC meet at a point <strong>and</strong> that this point bisects each ofthe lines.7.7 The edges OP, OQ <strong>and</strong> OR of a tetrahedron OPQR are vectors p, q <strong>and</strong> r,respectively, where p =2i +4j, q =2i − j +3k <strong>and</strong> r =4i − 2j +5k. Show thatOP is perpendicular to the plane containing OQR. Express the volume of thetetrahedron in terms of p, q <strong>and</strong> r <strong>and</strong> hence calculate the volume.7.8 Prove, by writing it out in component <strong>for</strong>m, that(a × b) × c =(a · c)b − (b · c)a,<strong>and</strong> deduce the result, stated in equation (7.25), that the operation of <strong>for</strong>mingthe vector product is non-associative.7.9 Prove Lagrange’s identity, i.e.(a × b) · (c × d) =(a · c)(b · d) − (a · d)(b · c).7.10 For four arbitrary vectors a, b, c <strong>and</strong> d, evaluate(a × b) × (c × d)in two different ways <strong>and</strong> so prove thata[b, c, d] − b[c, d, a]+c[d, a, b] − d[a, b, c] =0.Show that this reduces to the normal Cartesian representation of the vector d,i.e. d x i + d y j + d z k,ifa, b <strong>and</strong> c are taken as i, j <strong>and</strong> k, the Cartesian base vectors.7.11 Show that the points (1, 0, 1), (1, 1, 0) <strong>and</strong> (1, −3, 4) lie on a straight line. Give theequation of the line in the <strong>for</strong>mr = a + λb.235


VECTOR ALGEBRA7.12 The plane P 1 contains the points A, B <strong>and</strong> C, which have position vectorsa = −3i +2j, b =7i +2j <strong>and</strong> c =2i +3j +2k, respectively. Plane P 2 passesthrough A <strong>and</strong> is orthogonal to the line BC, whilst plane P 3 passes throughB <strong>and</strong> is orthogonal to the line AC. Find the coordinates of r, the point ofintersection of the three planes.7.13 Two planes have non-parallel unit normals ˆn <strong>and</strong> ˆm <strong>and</strong> their closest distancesfrom the origin are λ <strong>and</strong> µ, respectively. Find the vector equation of their lineof intersection in the <strong>for</strong>m r = νp + a.7.14 Two fixed points, A <strong>and</strong> B, in three-dimensional space have position vectors a<strong>and</strong> b. Identify the plane P given by(a − b) · r = 1 2 (a2 − b 2 ),where a <strong>and</strong> b are the magnitudes of a <strong>and</strong> b.Show also that the equation(a − r) · (b − r) =0describes a sphere S of radius |a − b|/2. Deduce that the intersection of P <strong>and</strong> Sis also the intersection of two spheres, centred on A <strong>and</strong> B, <strong>and</strong> each of radius|a − b|/ √ 2.7.15 Let O, A, B <strong>and</strong> C be four points with position vectors 0, a, b <strong>and</strong> c, <strong>and</strong> denoteby g = λa + µb + νc the position of the centre of the sphere on which they all lie.(a) Prove that λ, µ <strong>and</strong> ν simultaneously satisfy(a · a)λ +(a · b)µ +(a · c)ν = 1 2 a2<strong>and</strong> two other similar equations.(b) By making a change of origin, find the centre <strong>and</strong> radius of the sphere onwhich the points p =3i+j−2k, q =4i+3j−3k, r =7i−3k <strong>and</strong> s =6i+j−kall lie.7.16 The vectors a, b <strong>and</strong> c are coplanar <strong>and</strong> related byλa + µb + νc =0,where λ, µ, ν are not all zero. Show that the condition <strong>for</strong> the points with positionvectors αa, βb <strong>and</strong> γc to be collinear isλα + µ β + ν γ =0.7.17 Using vector methods:(a) Show that the line of intersection of the planes x +2y +3z = 0 <strong>and</strong>3x +2y + z = 0 is equally inclined to the x- <strong>and</strong>z-axes <strong>and</strong> makes an anglecos −1 (−2/ √ 6) with the y-axis.(b) Find the perpendicular distance between one corner of a unit cube <strong>and</strong> themajor diagonal not passing through it.7.18 Four points X i , i =1, 2, 3, 4, taken <strong>for</strong> simplicity as all lying within the octantx, y, z ≥ 0, have position vectors x i . Convince yourself that the direction ofvector x n lies within the sector of space defined by the directions of the otherthree vectors if[ ]xi · x jmin ,over j |x i ||x j |considered <strong>for</strong> i =1, 2, 3, 4 in turn, takes its maximum value <strong>for</strong> i = n, i.e.n equalsthat value of i <strong>for</strong> which the largest of the set of angles which x i makes withthe other vectors, is found to be the lowest. Determine whether any of the four236


7.10 EXERCISESabcdaFigure 7.17A face-centred cubic crystal.points with coordinatesX 1 =(3, 2, 2), X 2 =(2, 3, 1), X 3 =(2, 1, 3), X 4 =(3, 0, 3)lies within the tetrahedron defined by the origin <strong>and</strong> the other three points.7.19 The vectors a, b <strong>and</strong> c are not coplanar. The vectors a ′ , b ′ <strong>and</strong> c ′ are theassociated reciprocal vectors. Verify that the expressions (7.49)–(7.51) define a setof reciprocal vectors a ′ , b ′ <strong>and</strong> c ′ with the following properties:(a) a ′ · a = b ′ · b = c ′ · c =1;(b) a ′ · b = a ′ · c = b ′ · a etc = 0;(c) [a ′ , b ′ , c ′ ]=1/[a, b, c];(d) a =(b ′ × c ′ )/[a ′ , b ′ , c ′ ].7.20 Three non-coplanar vectors a, b <strong>and</strong> c, have as their respective reciprocal vectorsthe set a ′ , b ′ <strong>and</strong> c ′ . Show that the normal to the plane containing the pointsk −1 a, l −1 b <strong>and</strong> m −1 c is in the direction of the vector ka ′ + lb ′ + mc ′ .7.21 In a crystal with a face-centred cubic structure, the basic cell can be taken as acube of edge a with its centre at the origin of coordinates <strong>and</strong> its edges parallelto the Cartesian coordinate axes; atoms are sited at the eight corners <strong>and</strong> at thecentre of each face. However, other basic cells are possible. One is the rhomboidshown in figure 7.17, which has the three vectors b, c <strong>and</strong> d as edges.(a) Show that the volume of the rhomboid is one-quarter that of the cube.(b) Show that the angles between pairs of edges of the rhomboid are 60 ◦ <strong>and</strong> thatthe corresponding angles between pairs of edges of the rhomboid defined bythe reciprocal vectors to b, c, d are each 109.5 ◦ . (This rhomboid can be usedas the basic cell of a body-centred cubic structure, more easily visualised asa cube with an atom at each corner <strong>and</strong> one at its centre.)(c) In order to use the Bragg <strong>for</strong>mula, 2d sin θ = nλ, <strong>for</strong> the scattering of X-raysby a crystal, it is necessary to know the perpendicular distance d betweensuccessive planes of atoms; <strong>for</strong> a given crystal structure, d has a particularvalue <strong>for</strong> each set of planes considered. For the face-centred cubic structurefind the distance between successive planes with normals in the k, i + j <strong>and</strong>i + j + k directions.237


VECTOR ALGEBRA7.22 In subsection 7.6.2 we showed how the moment or torque of a <strong>for</strong>ce about an axiscould be represented by a vector in the direction of the axis. The magnitude ofthe vector gives the size of the moment <strong>and</strong> the sign of the vector gives the sense.Similar representations can be used <strong>for</strong> angular velocities <strong>and</strong> angular momenta.(a) The magnitude of the angular momentum about the origin of a particle ofmass m moving with velocity v on a path that is a perpendicular distance dfrom the origin is given by m|v|d. Show that if r is the position of the particlethen the vector J = r × mv represents the angular momentum.(b) Now consider a rigid collection of particles (or a solid body) rotating aboutan axis through the origin, the angular velocity of the collection beingrepresented by ω.(i) Show that the velocity of the ith particle isv i = ω × r i<strong>and</strong> that the total angular momentum J isJ = ∑ im i [r 2 i ω − (r i · ω)r i ].(ii) Show further that the component of J along the axis of rotation canbe written as Iω, whereI, the moment of inertia of the collectionabout the axis or rotation, is given byI = ∑ im i ρ 2 i .Interpret ρ i geometrically.(iii) Prove that the total kinetic energy of the particles is 1 2 Iω2 .7.23 By proceeding as indicated below, prove the parallel axis theorem, whichstatesthat, <strong>for</strong> a body of mass M, the moment of inertia I about any axis is related tothe corresponding moment of inertia I 0 about a parallel axis that passes throughthe centre of mass of the body byI = I 0 + Ma 2 ⊥,where a ⊥ is the perpendicular distance between the two axes. Note that I 0 canbe written as∫(ˆn × r) · (ˆn × r) dm,where r is the vector position, relative to the centre of mass, of the infinitesimalmass dm <strong>and</strong> ˆn is a unit vector in the direction of the axis of rotation. Write asimilar expression <strong>for</strong> I in which r is replaced by r ′ = r − a, wherea is the vectorposition of any point on the axis to which I refers. Use Lagrange’s identity <strong>and</strong>the fact that ∫ r dm = 0 (by the definition of the centre of mass) to establish theresult.7.24 Without carrying out any further integration, use the results of the previousexercise, the worked example in subsection 6.3.4 <strong>and</strong> exercise 6.10 to prove thatthe moment of inertia of a uni<strong>for</strong>m rectangular lamina, of mass M <strong>and</strong> sides a<strong>and</strong> b, about an axis perpendicular to its plane <strong>and</strong> passing through the point(αa/2,βb/2), with −1 ≤ α, β ≤ 1, isM12 [a2 (1 + 3α 2 )+b 2 (1 + 3β 2 )].238


7.10 EXERCISESV 0 cos ωtV 4V 1V 2R 1 =50ΩI 2I 1I 3LC =10µFV 3R 2Figure 7.18 An oscillatory electric circuit. The power supply has angularfrequency ω =2πf = 400π s −1 .7.25 Define a set of (non-orthogonal) base vectors a = j + k, b = i + k <strong>and</strong> c = i + j.(a) Establish their reciprocal vectors <strong>and</strong> hence express the vectors p =3i−2j+k,q = i +4j <strong>and</strong> r = −2i + j + k in terms of the base vectors a, b <strong>and</strong> c.(b) Verify that the scalar product p · q has the same value, −5, when evaluatedusing either set of components.7.26 Systems that can be modelled as damped harmonic oscillators are widespread;pendulum clocks, car shock absorbers, tuning circuits in television sets <strong>and</strong> radios,<strong>and</strong> collective electron motions in plasmas <strong>and</strong> metals are just a few examples.In all these cases, one or more variables describing the system obey(s) anequation of the <strong>for</strong>mẍ +2γẋ + ω0x 2 = P cos ωt,where ẋ = dx/dt, etc. <strong>and</strong> the inclusion of the factor 2 is conventional. In thesteady state (i.e. after the effects of any initial displacement or velocity have beendamped out) the solution of the equation takes the <strong>for</strong>mx(t) =A cos(ωt + φ).By expressing each term in the <strong>for</strong>m B cos(ωt+ɛ), <strong>and</strong> representing it by a vectorof magnitude B making an angle ɛ with the x-axis, draw a closed vector diagram,at t = 0, say, that is equivalent to the equation.(a) Convince yourself that whatever the value of ω (> 0) φ must be negative(−π


VECTOR ALGEBRAof vector plots <strong>for</strong> potential differences <strong>and</strong> currents (they could all be on thesame plot if suitable scales were chosen), determine all unknown currents <strong>and</strong>potential differences <strong>and</strong> find values <strong>for</strong> the inductance of L <strong>and</strong> the resistanceof R 2 .[ Scales of 1 cm = 0.1 V <strong>for</strong> potential differences <strong>and</strong> 1 cm = 1 mA <strong>for</strong> currentsare convenient. ]7.11 Hints <strong>and</strong> answers7.1 (c), (d) <strong>and</strong> (e).7.3 (a) A sphere of radius k centred on the origin; (b) a plane with its normal in thedirection of u <strong>and</strong> at a distance l from the origin; (c) a cone with its axis parallelto u <strong>and</strong> of semiangle cos −1 m; (d) a circular cylinder of radius n with its axisparallel to u.7.5 (a) cos √ −1 2/3; (b) z − x =2;(c)1/ √ 2; (d) 1 1[ (c × g) · j = 1 .3 2 37.7 Show that q × r is parallel to p; volume = 1 1 (q × r) · p] = 5 .3 2 37.9 Note that (a × b) · (c × d) =d · [(a × b) × c] <strong>and</strong> use the result <strong>for</strong> a triple vectorproduct to exp<strong>and</strong> the expression in square brackets.7.11 Show that the position vectors of the points are linearly dependent; r = a + λbwhere a = i + k <strong>and</strong> b = −j + k.7.13 Show that p must have the direction ˆn× ˆm <strong>and</strong> write a as xˆn+y ˆm. By obtaining apair of simultaneous equations <strong>for</strong> x <strong>and</strong> y, prove that x =(λ−µˆn· ˆm)/[1−(ˆn· ˆm) 2 ]<strong>and</strong> that y =(µ − λˆn · ˆm)/[1 − (ˆn · ˆm) 2 ].7.15 (a) Note that |a − g| 2 = R 2 = |0 − g| 2 , leading to a · a =2a · g.(b) Make p the new origin <strong>and</strong> solve the three simultaneous linear equations toobtain λ =5/18, µ =10/18, ν = −3/18, giving g =2i − k <strong>and</strong> a sphere ofradius √ 5 centred on (5, 1, −3).7.17 (a) Find two points on both planes, say (0, 0, 0) <strong>and</strong> (1, −2, 1), <strong>and</strong> hence determinethe direction cosines of the line of intersection; (b) ( 2 3 )1/2 .7.19 For (c) <strong>and</strong> (d), treat (c × a) × (a × b) as a triple vector product with c × a as oneof the three vectors.7.21 (b) b ′ = a −1 (−i + j + k), c ′ = a −1 (i − j + k), d ′ = a −1 (i + j − k); (c) a/2 <strong>for</strong>directionk; successive planes through (0, 0, 0) <strong>and</strong> (a/2, 0,a/2) give a spacing of a/ √ 8<strong>for</strong>direction i + j; successive planes through (−a/2, 0, 0) <strong>and</strong> (a/2, 0, 0) give a spacingof a/ √ 3<strong>for</strong>directioni + j + k.7.23 Note that a 2 − (ˆn · a) 2 = a 2 ⊥ .7.25 p = −2a +3b, q = 3 a − 3 b + 5 c <strong>and</strong> r =2a − b − c. Remember that a · a = b · b =2 2 2c · c =2<strong>and</strong>a · b = a · c = b · c =1.7.27 With currents in mA <strong>and</strong> potential differences in volts:I 1 =(7.76, −23.2 ◦ ), I 2 =(14.36, −50.8 ◦ ), I 3 =(8.30, 103.4 ◦ );V 1 =(0.388, −23.2 ◦ ), V 2 =(0.287, −50.8 ◦ ), V 4 =(0.596, 39.2 ◦ );L = 33 mH, R 2 =20Ω.240


8Matrices <strong>and</strong> vector spacesIn the previous chapter we defined a vector as a geometrical object which hasboth a magnitude <strong>and</strong> a direction <strong>and</strong> which may be thought of as an arrow fixedin our familiar three-dimensional space, a space which, if we need to, we defineby reference to, say, the fixed stars. This geometrical definition of a vector is bothuseful <strong>and</strong> important since it is independent of any coordinate system with whichwe choose to label points in space.In most specific applications, however, it is necessary at some stage to choosea coordinate system <strong>and</strong> to break down a vector into its component vectors inthe directions of increasing coordinate values. Thus <strong>for</strong> a particular Cartesiancoordinate system (<strong>for</strong> example) the component vectors of a vector a will be a x i,a y j <strong>and</strong> a z k <strong>and</strong> the complete vector will bea = a x i + a y j + a z k. (8.1)Although we have so far considered only real three-dimensional space, we mayextend our notion of a vector to more abstract spaces, which in general canhave an arbitrary number of dimensions N. We may still think of such a vectoras an ‘arrow’ in this abstract space, so that it is again independent of any (Ndimensional)coordinate system with which we choose to label the space. As anexample of such a space, which, though abstract, has very practical applications,we may consider the description of a mechanical or electrical system. If the stateof a system is uniquely specified by assigning values to a set of N variables,which could be angles or currents, <strong>for</strong> example, then that state can be representedby a vector in an N-dimensional space, the vector having those values as itscomponents.In this chapter we first discuss general vector spaces <strong>and</strong> their properties. Wethen go on to discuss the trans<strong>for</strong>mation of one vector into another by a linearoperator. This leads naturally to the concept of a matrix, a two-dimensional arrayof numbers. The properties of matrices are then discussed <strong>and</strong> we conclude with241


MATRICES AND VECTOR SPACESa discussion of how to use these properties to solve systems of linear equations.The application of matrices to the study of oscillations in physical systems istakenupinchapter9.8.1 Vector spacesA set of objects (vectors) a, b, c, ... is said to <strong>for</strong>m a linear vector space V if:(i) the set is closed under commutative <strong>and</strong> associative addition, so thata + b = b + a, (8.2)(a + b)+c = a +(b + c); (8.3)(ii) the set is closed under multiplication by a scalar (any complex number) to<strong>for</strong>m a new vector λa, the operation being both distributive <strong>and</strong> associativeso thatλ(a + b) =λa + λb, (8.4)(λ + µ)a = λa + µa, (8.5)λ(µa) =(λµ)a, (8.6)where λ <strong>and</strong> µ are arbitrary scalars;(iii) there exists a null vector 0 such that a + 0 = a <strong>for</strong> all a;(iv) multiplication by unity leaves any vector unchanged, i.e. 1 × a = a;(v) all vectors have a corresponding negative vector −a such that a+(−a) =0.It follows from (8.5) with λ = 1 <strong>and</strong> µ = −1 that−a is the same vector as(−1) × a.We note that if we restrict all scalars to be real then we obtain a real vectorspace (an example of which is our familiar three-dimensional space); otherwise,in general, we obtain a complex vector space. We note that it is common to use theterms ‘vector space’ <strong>and</strong> ‘space’, instead of the more <strong>for</strong>mal ‘linear vector space’.The span of a set of vectors a, b,...,s is defined as the set of all vectors thatmay be written as a linear sum of the original set, i.e. all vectorsx = αa + βb + ···+ σs (8.7)that result from the infinite number of possible values of the (in general complex)scalars α,β,...,σ.Ifx in (8.7) is equal to 0 <strong>for</strong> some choice of α,β,...,σ (not allzero), i.e. ifαa + βb + ···+ σs = 0, (8.8)then the set of vectors a, b,...,s, issaidtobelinearly dependent. Insuchasetat least one vector is redundant, since it can be expressed as a linear sum ofthe others. If, however, (8.8) is not satisfied by any set of coefficients (other than242


8.1 VECTOR SPACESthe trivial case in which all the coefficients are zero) then the vectors are linearlyindependent, <strong>and</strong> no vector in the set can be expressed as a linear sum of theothers.If, in a given vector space, there exist sets of N linearly independent vectors,but no set of N + 1 linearly independent vectors, then the vector space is said tobe N-dimensional. (In this chapter we will limit our discussion to vector spaces offinite dimensionality; spaces of infinite dimensionality are discussed in chapter 17.)8.1.1 Basis vectorsIf V is an N-dimensional vector space then any set of N linearly independentvectors e 1 , e 2 ,...,e N <strong>for</strong>ms a basis <strong>for</strong> V .Ifx is an arbitrary vector lying in V thenthe set of N + 1 vectors x, e 1 , e 2 ,...,e N , must be linearly dependent <strong>and</strong> there<strong>for</strong>esuch thatαe 1 + βe 2 + ···+ σe N + χx = 0, (8.9)where the coefficients α,β,...,χ are not all equal to 0, <strong>and</strong> in particular χ ≠0.Rearranging (8.9) we may write x as a linear sum of the vectors e i as follows:x = x 1 e 1 + x 2 e 2 + ···+ x N e N =N∑x i e i , (8.10)<strong>for</strong> some set of coefficients x i that are simply related to the original coefficients,e.g. x 1 = −α/χ, x 2 = −β/χ, etc. Since any x lying in the span of V can beexpressed in terms of the basis or base vectors e i , the latter are said to <strong>for</strong>ma complete set. The coefficients x i are the components of x with respect to thee i -basis. These components are unique, since if boththeni=1i=1N∑N∑x = x i e i <strong>and</strong> x = y i e i ,i=1N∑(x i − y i )e i = 0, (8.11)i=1which, since the e i are linearly independent, has only the solution x i = y i <strong>for</strong> alli =1, 2,...,N.From the above discussion we see that any set of N linearly independentvectors can <strong>for</strong>m a basis <strong>for</strong> an N-dimensional space. If we choose a different sete ′ i , i =1,...,N then we can write x asx = x ′ 1e ′ 1 + x ′ 2e ′ 2 + ···+ x ′ Ne ′ N =243N∑x ′ ie ′ i. (8.12)i=1


MATRICES AND VECTOR SPACESWe reiterate that the vector x (a geometrical entity) is independent of the basis– it is only the components of x that depend on the basis. We note, however,that given a set of vectors u 1 , u 2 ,...,u M ,whereM ≠ N, inanN-dimensionalvector space, then either there exists a vector that cannot be expressed as alinear combination of the u i or, <strong>for</strong> some vector that can be so expressed, thecomponents are not unique.8.1.2 The inner productWe may usefully add to the description of vectors in a vector space by definingthe inner product of two vectors, denoted in general by 〈a|b〉, which is a scalarfunction of a <strong>and</strong> b. The scalar or dot product, a · b ≡|a||b| cos θ, ofvectorsin real three-dimensional space (where θ is the angle between the vectors), wasintroduced in the last chapter <strong>and</strong> is an example of an inner product. In effect thenotion of an inner product 〈a|b〉 is a generalisation of the dot product to moreabstract vector spaces. Alternative notations <strong>for</strong> 〈a|b〉 are (a, b), or simply a · b.The inner product has the following properties:(i) 〈a|b〉 = 〈b|a〉 ∗ ,(ii) 〈a|λb + µc〉 = λ〈a|b〉 + µ〈a|c〉.We note that in general, <strong>for</strong> a complex vector space, (i) <strong>and</strong> (ii) imply that〈λa + µb|c〉 = λ ∗ 〈a|c〉 + µ ∗ 〈b|c〉, (8.13)〈λa|µb〉 = λ ∗ µ〈a|b〉. (8.14)Following the analogy with the dot product in three-dimensional real space,two vectors in a general vector space are defined to be orthogonal if 〈a|b〉 =0.Similarly, the norm of a vector a is given by ‖a‖ = 〈a|a〉 1/2 <strong>and</strong> is clearly ageneralisation of the length or modulus |a| of a vector a in three-dimensionalspace. In a general vector space 〈a|a〉 can be positive or negative; however, weshall be primarily concerned with spaces in which 〈a|a〉 ≥0 <strong>and</strong> which are thussaid to have a positive semi-definite norm. Insuchaspace〈a|a〉 = 0 implies a = 0.Let us now introduce into our N-dimensional vector space a basis ê 1 , ê 2 ,...,ê Nthat has the desirable property of being orthonormal (the basis vectors are mutuallyorthogonal <strong>and</strong> each has unit norm), i.e. a basis that has the property〈ê i |ê j 〉 = δ ij . (8.15)Here δ ij is the Kronecker delta symbol (of which we say more in chapter 26) <strong>and</strong>has the properties{1 <strong>for</strong> i = j,δ ij =0 <strong>for</strong> i ≠ j.244


8.1 VECTOR SPACESIn the above basis we may express any two vectors a <strong>and</strong> b asN∑N∑a = a i ê i <strong>and</strong> b = b i ê i .i=1Furthermore, in such an orthonormal basis we have, <strong>for</strong> any a,N∑N∑〈ê j |a〉 = 〈ê j |a i ê i 〉 = a i 〈ê j |ê i 〉 = a j . (8.16)i=1Thus the components of a are given by a i = 〈ê i |a〉. Note that this is not trueunless the basis is orthonormal. We can write the inner product of a <strong>and</strong> b interms of their components in an orthonormal basis as〈a|b〉 = 〈a 1 ê 1 + a 2 ê 2 + ···+ a N ê N |b 1 ê 1 + b 2 ê 2 + ···+ b N ê N 〉N∑N∑ N∑= a ∗ i b i 〈ê i |ê i 〉 + a ∗ i b j 〈ê i |ê j 〉=i=1N∑a ∗ i b i,i=1i=1j≠iwhere the second equality follows from (8.14) <strong>and</strong> the third from (8.15). This isclearly a generalisation of the expression (7.21) <strong>for</strong> the dot product of vectors inthree-dimensional space.We may generalise the above to the case where the base vectors e 1 , e 2 ,...,e Nare not orthonormal (or orthogonal). In general we can define the N 2 numbersi=1i=1G ij = 〈e i |e j 〉. (8.17)Then, if a = ∑ Ni=1 a ie i <strong>and</strong> b = ∑ Ni=1 b ie i , the inner product of a <strong>and</strong> b is given by〈 ∣∑ N ∣∣∣∣∣ 〉∑N〈a|b〉 = a i e i b j e j==N∑i=1i=1 j=1N∑i=1 j=1j=1N∑a ∗ i b j 〈e i |e j 〉N∑a ∗ i G ij b j . (8.18)We further note that from (8.17) <strong>and</strong> the properties of the inner product werequire G ij = G ∗ ji . This in turn ensures that ‖a‖ = 〈a|a〉 is real, since thenN∑ N∑N∑ N∑〈a|a〉 ∗ = a i G ∗ ija ∗ j = a ∗ j G ji a i = 〈a|a〉.i=1 j=1j=1 i=1245


MATRICES AND VECTOR SPACES8.1.3 Some useful inequalitiesFor a set of objects (vectors) <strong>for</strong>ming a linear vector space in which 〈a|a〉 ≥0<strong>for</strong>all a, the following inequalities are often useful.(i) Schwarz’s inequality is the most basic result <strong>and</strong> states that|〈a|b〉| ≤ ‖a‖‖b‖, (8.19)where the equality holds when a is a scalar multiple of b, i.e.whena = λb.It is important here to distinguish between the absolute value of a scalar,|λ|, <strong>and</strong>thenorm of a vector, ‖a‖. Schwarz’s inequality may be proved byconsidering‖a + λb‖ 2 = 〈a + λb|a + λb〉= 〈a|a〉 + λ〈a|b〉 + λ ∗ 〈b|a〉 + λλ ∗ 〈b|b〉.If we write 〈a|b〉 as |〈a|b〉|e iα then‖a + λb‖ 2 = ‖a‖ 2 + |λ| 2 ‖b‖ 2 + λ|〈a|b〉|e iα + λ ∗ |〈a|b〉|e −iα .However, ‖a + λb‖ 2 ≥ 0 <strong>for</strong> all λ, sowemaychooseλ = re −iα <strong>and</strong> requirethat, <strong>for</strong> all r,0 ≤‖a + λb‖ 2 = ‖a‖ 2 + r 2 ‖b‖ 2 +2r|〈a|b〉|.This means that the quadratic equation in r <strong>for</strong>med by setting the RHSequal to zero must have no real roots. This, in turn, implies that4|〈a|b〉| 2 ≤ 4‖a‖ 2 ‖b‖ 2 ,which, on taking the square root (all factors are necessarily positive) ofboth sides, gives Schwarz’s inequality.(ii) The triangle inequality states that‖a + b‖ ≤‖a‖ + ‖b‖ (8.20)<strong>and</strong> may be derived from the properties of the inner product <strong>and</strong> Schwarz’sinequality as follows. Let us first consider‖a + b‖ 2 = ‖a‖ 2 + ‖b‖ 2 +2Re〈a|b〉 ≤‖a‖ 2 + ‖b‖ 2 +2|〈a|b〉|.Using Schwarz’s inequality we then have‖a + b‖ 2 ≤‖a‖ 2 + ‖b‖ 2 +2‖a‖‖b‖ =(‖a‖ + ‖b‖) 2 ,which, on taking the square root, gives the triangle inequality (8.20).(iii) Bessel’s inequality requires the introduction of an orthonormal basis ê i ,i =1, 2,...,N into the N-dimensional vector space; it states that‖a‖ 2 ≥ ∑ i|〈ê i |a〉| 2 , (8.21)246


8.2 LINEAR OPERATORSwhere the equality holds if the sum includes all N basis vectors. If notall the basis vectors are included in the sum then the inequality results(though of course the equality remains if those basis vectors omitted allhave a i = 0). Bessel’s inequality can also be written〈a|a〉 ≥ ∑ i|a i | 2 ,where the a i are the components of a in the orthonormal basis. From (8.16)these are given by a i = 〈ê i |a〉. The above may be proved by considering∣ a − ∑ i∣ ∣∣∣ ∣∣∣ 〈〈ê i |a〉ê i2= a − ∑ i∣ ∣∣a ∑〈ê i |a〉ê i − 〈ê j |a〉ê j〉.Exp<strong>and</strong>ing out the inner product <strong>and</strong> using 〈ê i |a〉 ∗ = 〈a|ê i 〉, we obtain∣ a − ∑ ∣ ∣∣∣ ∣∣∣2〈ê i |a〉ê i = 〈a|a〉−2 ∑ 〈a|ê i 〉〈ê i |a〉 + ∑ ∑〈a|ê i 〉〈ê j |a〉〈ê i |ê j 〉.iii jNow 〈ê i |ê j 〉 = δ ij , since the basis is orthonormal, <strong>and</strong> so we find0 ≤∣ a − ∑ ∣ ∣∣∣ ∣∣∣2〈ê i |a〉ê i = ‖a‖ 2 − ∑ |〈ê i |a〉| 2 ,iiwhich is Bessel’s inequality.We take this opportunity to mention also(iv) the parallelogram equality‖a + b‖ 2 + ‖a − b‖ 2 =2 ( ‖a‖ 2 + ‖b‖ 2) , (8.22)which may be proved straight<strong>for</strong>wardly from the properties of the innerproduct.j8.2 Linear operatorsWe now discuss the action of linear operators on vectors in a vector space. Alinear operator A associates with every vector x another vectory = A x,in such a way that, <strong>for</strong> two vectors a <strong>and</strong> b,A (λa + µb) =λA a + µA b,where λ, µ are scalars. We say that A ‘operates’ on x to give the vector y. Wenote that the action of A is independent of any basis or coordinate system <strong>and</strong>247


MATRICES AND VECTOR SPACESmay be thought of as ‘trans<strong>for</strong>ming’ one geometrical entity (i.e. a vector) intoanother.If we now introduce a basis e i , i =1, 2,...,N, into our vector space then theaction of A on each of the basis vectors is to produce a linear combination ofthe latter; this may be written asA e j =N∑A ij e i , (8.23)i=1where A ij is the ith component of the vector A e j in this basis; collectively thenumbers A ij are called the components of the linear operator in the e i -basis. Inthis basis we can express the relation y = A x in component <strong>for</strong>m asy =⎛ ⎞N∑N∑y i e i = A ⎝ x j e j⎠ =i=1<strong>and</strong> hence, in purely component <strong>for</strong>m, in this basis we havey i =j=1N∑j=1x jN ∑i=1A ij e i ,N∑A ij x j . (8.24)j=1If we had chosen a different basis e ′ i , in which the components of x, y <strong>and</strong> Aare x ′ i , y′ i <strong>and</strong> A ′ ij respectively then the geometrical relationship y = A x would berepresented in this new basis byy ′ i =N∑A ′ ijx ′ j.j=1We have so far assumed that the vector y is in the same vector space asx. If,however,y belongs to a different vector space, which may in general beM-dimensional (M ≠ N) then the above analysis needs a slight modification. Byintroducing a basis set f i , i =1, 2,...,M, into the vector space to which y belongswe may generalise (8.23) asA e j =M∑A ij f i ,i=1where the components A ij of the linear operator A relate to both of the bases e j<strong>and</strong> f i .248


8.3 MATRICES8.2.1 Properties of linear operatorsIf x is a vector <strong>and</strong> A <strong>and</strong> B are two linear operators then it follows that(A + B )x = A x + B x,(λA )x = λ(A x),(AB)x = A (B x),where in the last equality we see that the action of two linear operators insuccession is associative. The product of two linear operators is not in generalcommutative, however, so that in general ABx ≠ BAx. In an obvious way wedefine the null (or zero) <strong>and</strong> identity operators byO x = 0 <strong>and</strong> I x = x,<strong>for</strong> any vector x in our vector space. Two operators A <strong>and</strong> B are equal ifA x = B x <strong>for</strong> all vectors x. Finally, if there exists an operator A −1 such thatAA −1 = A −1 A = Ithen A −1 is the inverse of A . Some linear operators do not possess an inverse<strong>and</strong> are called singular, whilst those operators that do have an inverse are termednon-singular.8.3 MatricesWe have seen that in a particular basis e i both vectors <strong>and</strong> linear operatorscan be described in terms of their components with respect to the basis. Thesecomponents may be displayed as an array of numbers called a matrix. Ingeneral,if a linear operator A trans<strong>for</strong>ms vectors from an N-dimensional vector space,<strong>for</strong> which we choose a basis e j , j =1, 2,...,N, into vectors belonging to anM-dimensional vector space, with basis f i , i =1, 2,...,M, then we may representthe operator A by the matrix⎛⎞A 11 A 12 ... A 1NA 21 A 22 ... A 2NA = ⎜⎝. ⎟. . .. . ⎠ . (8.25)A M1 A M2 ... A MNThe matrix elements A ij are the components of the linear operator with respectto the bases e j <strong>and</strong> f i ; the component A ij of the linear operator appears in theith row <strong>and</strong> jth column of the matrix. The array has M rows <strong>and</strong> N columns<strong>and</strong> is thus called an M × N matrix. If the dimensions of the two vector spacesare the same, i.e. M = N (<strong>for</strong> example, if they are the same vector space) then wemay represent A by an N × N or square matrix of order N. The component A ij ,which in general may be complex, is also denoted by (A) ij .249


MATRICES AND VECTOR SPACESIn a similar way we may denote a vector x in terms of its components x i in abasis e i , i =1, 2,...,N, by the array⎛ ⎞x 1x 2x = ⎜ ⎟⎝. ⎠ ,x Nwhich is a special case of (8.25) <strong>and</strong> is called a column matrix (or conventionally,<strong>and</strong> slightly confusingly, a column vector or even just a vector – strictly speakingthe term ‘vector’ refers to the geometrical entity x). The column matrix x can alsobe written asx =(x 1 x 2 ··· x N ) T ,which is the transpose of a row matrix (see section 8.6).We note that in a different basis e ′ i the vector x would be represented by adifferent column matrix containing the components x ′ i in the new basis, i.e.⎛x ′ = ⎜⎝Thus, we use x <strong>and</strong> x ′ to denote different column matrices which, in different basese i <strong>and</strong> e ′ i , represent the same vector x. In many texts, however, this distinction isnot made <strong>and</strong> x (rather than x) is equated to the corresponding column matrix; ifwe regard x as the geometrical entity, however, this can be misleading <strong>and</strong> so weexplicitly make the distinction. A similar argument follows <strong>for</strong> linear operators;the same linear operator A is described in different bases by different matrices A<strong>and</strong> A ′ , containing different matrix elements.x ′ 1x ′ 2..x ′ N⎞⎟⎠ .8.4 Basic matrix algebraThe basic algebra of matrices may be deduced from the properties of the linearoperators that they represent. In a given basis the action of two linear operatorsA <strong>and</strong> B on an arbitrary vector x (see the beginning of subsection 8.2.1), whenwritten in terms of components using (8.24), is given by∑(A + B) ij x j = ∑ A ij x j + ∑ B ij x j ,jjj∑(λA) ij x j = λ ∑ A ij x j ,jj∑(AB) ij x j = ∑ A ik (Bx) k = ∑ ∑A ik B kj x j .jkj k250


8.4 BASIC MATRIX ALGEBRANow, since x is arbitrary, we can immediately deduce the way in which matricesare added or multiplied, i.e.(A + B) ij = A ij + B ij , (8.26)(λA) ij = λA ij , (8.27)(AB) ij = ∑ A ik B kj .k(8.28)We note that a matrix element may, in general, be complex. We now discussmatrix addition <strong>and</strong> multiplication in more detail.8.4.1 Matrix addition <strong>and</strong> multiplication by a scalarFrom (8.26) we see that the sum of two matrices, S = A + B, is the matrix whoseelements are given byS ij = A ij + B ij<strong>for</strong> every pair of subscripts i, j, with i =1, 2,...,M <strong>and</strong> j =1, 2,...,N. Forexample, if A <strong>and</strong> B are 2 × 3 matrices then S = A + B is given by( ) ( )S11 S 12 S 13 A11 A=12 A 13+S 21 S 22 S 23 A 21 A 22 A 23=( )B11 B 12 B 13B 21 B 22 B 23( )A11 + B 11 A 12 + B 12 A 13 + B 13. (8.29)A 21 + B 21 A 22 + B 22 A 23 + B 23Clearly, <strong>for</strong> the sum of two matrices to have any meaning, the matrices must havethe same dimensions, i.e. both be M × N matrices.From definition (8.29) it follows that A + B = B + A <strong>and</strong> that the sum of anumber of matrices can be written unambiguously without bracketting, i.e. matrixaddition is commutative <strong>and</strong> associative.The difference of two matrices is defined by direct analogy with addition. Thematrix D = A − B has elementsD ij = A ij − B ij , <strong>for</strong> i =1, 2,...,M, j =1, 2,...,N. (8.30)From (8.27) the product of a matrix A with a scalar λ is the matrix withelements λA ij , <strong>for</strong> example( ) ( )A11 Aλ12 A 13 λA11 λA=12 λA 13. (8.31)A 21 A 22 A 23 λA 21 λA 22 λA 23Multiplication by a scalar is distributive <strong>and</strong> associative.251


MATRICES AND VECTOR SPACES◮The matrices A, B <strong>and</strong> C are given by( ) ( ) ( )2 −11 0−2 1A =, B =, C =.3 10 −2−1 1Find the matrix D = A +2B − C.(2 −1D =3 1=) (1 0+20 −2) (−2 1−−1 1(2+2× 1 − (−2) −1+2× 0 − 13+2× 0 − (−1) 1+2× (−2) − 1))=( )6 −2. ◭4 −4From the above considerations we see that the set of all, in general complex,M × N matrices (with fixed M <strong>and</strong> N) <strong>for</strong>ms a linear vector space of dimensionMN. One basis <strong>for</strong> the space is the set of M × N matrices E (p,q) with the propertythat E (p,q)ij =1ifi = p <strong>and</strong> j = q whilst E (p,q)ij = 0 <strong>for</strong> all other values of i <strong>and</strong>j, i.e. each matrix has only one non-zero entry, which equals unity. Here the pair(p, q) is simply a label that picks out a particular one of the matrices E (p,q) ,thetotal number of which is MN.8.4.2 Multiplication of matricesLet us consider again the ‘trans<strong>for</strong>mation’ of one vector into another, y = A x,which, from (8.24), may be described in terms of components with respect to aparticular basis asy i =N∑A ij x j <strong>for</strong> i =1, 2,...,M. (8.32)j=1Writing this in matrix <strong>for</strong>m as y = Ax we have⎛ ⎞⎛⎛⎞y 1 A 11 A 12 ... A 1Ny 2⎜ ⎟⎝ . ⎠ = A 21 A 22 ... A 2N⎜⎝.. ⎟. .. . ⎠ ⎜⎝A M1 A M2 ... A MNy Mx 1x 2..x N⎞⎟⎠(8.33)where we have highlighted with boxes the components used to calculate theelement y 2 : using (8.32) <strong>for</strong> i =2,y 2 = A 21 x 1 + A 22 x 2 + ···+ A 2N x N .All the other components y i are calculated similarly.If instead we operate with A on a basis vector e j having all components zero252


8.4 BASIC MATRIX ALGEBRAexcept <strong>for</strong> the jth, which equals unity, then we find⎛ ⎞0⎛⎞A 11 A 12 ... A 1N0⎛A 21 A 22 ... A 2N..Ae j = ⎜⎝. ⎟= ⎜. . .. . ⎠1⎝A M1 A M2 ... A⎜ . ⎟MN ⎝ . ⎠0<strong>and</strong> so confirm our identification of the matrix element A ij as the ith componentof Ae j in this basis.From (8.28) we can extend our discussion to the product of two matricesP = AB, whereP is the matrix of the quantities <strong>for</strong>med by the operation ofthe rows of A on the columns of B, treating each column of B in turn as thevector x represented in component <strong>for</strong>m in (8.32). It is clear that, <strong>for</strong> this to bea meaningful definition, the number of columns in A must equal the number ofrows in B. Thus the product AB of an M × N matrix A with an N × R matrix Bis itself an M × R matrix P, whereN∑P ij = A ik B kj <strong>for</strong> i =1, 2,...,M, j =1, 2,...,R.k=1For example, P = AB may be written in matrix <strong>for</strong>m() ( ) ⎛ P 11 P 12 A11 A=12 A 13 ⎝P 21 P 22 A 21 A 22 A 23whereP 11 = A 11 B 11 + A 12 B 21 + A 13 B 31 ,P 21 = A 21 B 11 + A 22 B 21 + A 23 B 31 ,P 12 = A 11 B 12 + A 12 B 22 + A 13 B 32 ,P 22 = A 21 B 12 + A 22 B 22 + A 23 B 32 .A 1jA 2j.A Mj⎞⎟⎠ ,B 11 B 12⎞B 21 B 22⎠Multiplication of more than two matrices follows naturally <strong>and</strong> is associative.So, <strong>for</strong> example,A(BC) ≡ (AB)C, (8.34)provided, of course, that all the products are defined.As mentioned above, if A is an M × N matrix <strong>and</strong> B is an N × M matrix thentwo product matrices are possible, i.e.P = AB <strong>and</strong> Q = BA.253


MATRICES AND VECTOR SPACESThese are clearly not the same, since P is an M × M matrix whilst Q is anN × N matrix. Thus, particular care must be taken to write matrix products inthe intended order; P = AB but Q = BA. We note in passing that A 2 means AA,A 3 means A(AA) =(AA)A etc. Even if both A <strong>and</strong> B are square, in generalAB ≠ BA, (8.35)i.e. the multiplication of matrices is not, in general, commutative.◮Evaluate P = AB <strong>and</strong> Q = BA where⎛A = ⎝ 3 2 −10 3 21 −3 4⎞ ⎛⎠ , B =⎝ 2 −2 31 1 03 2 1⎞⎠ .As we saw <strong>for</strong> the 2 × 2 case above, the element P ij of the matrix P = AB is found bymentally taking the ‘scalar product’ of the ith row of A with the jth column of B. Forexample, P 11 =3× 2+2× 1+(−1) × 3=5,P 12 =3× (−2) + 2 × 1+(−1) × 2=−6, etc.Thus⎛P = AB = ⎝ 3 2 −1⎞ ⎛0 3 2 ⎠ ⎝ 2 −2 3⎞ ⎛1 1 0 ⎠ = ⎝ 5 −6 8⎞9 7 2 ⎠ ,1 −3 4 3 2 1 11 3 7<strong>and</strong>, similarly,⎛Q = BA = ⎝ 2 −2 3⎞ ⎛1 1 0 ⎠ ⎝ 3 2 −1⎞ ⎛0 3 2 ⎠ = ⎝ 9 −11 6⎞3 5 1 ⎠ .3 2 1 1 −3 4 10 9 5These results illustrate that, in general, two matrices do not commute. ◭<strong>and</strong>The property that matrix multiplication is distributive over addition, i.e. thatfollows directly from its definition.(A + B)C = AC + BC (8.36)C(A + B) =CA + CB, (8.37)8.4.3 The null <strong>and</strong> identity matricesBoth the null matrix <strong>and</strong> the identity matrix are frequently encountered, <strong>and</strong> wetake this opportunity to introduce them briefly, leaving their uses until later. Thenull or zero matrix 0 has all elements equal to zero, <strong>and</strong> so its properties areA0 = 0 = 0A,A + 0 = 0 + A = A.254


8.5 FUNCTIONS OF MATRICESThe identity matrix I has the propertyAI = IA = A.It is clear that, in order <strong>for</strong> the above products to be defined, the identity matrixmust be square. The N × N identity matrix (often denoted by I N )hasthe<strong>for</strong>m⎛⎞1 0 ··· 0.I N =0 1 .⎜ .⎝. ⎟. .. 0 ⎠ .0 ··· 0 18.5 Functions of matricesIf a matrix A is square then, as mentioned above, one can define powers of A ina straight<strong>for</strong>ward way. For example A 2 = AA, A 3 = AAA, or in the general caseA n = AA ···A(n times),where n is a positive integer. Having defined powers of a square matrix A, wemay construct functions of A of the <strong>for</strong>mS = ∑ na n A n ,where the a k are simple scalars <strong>and</strong> the number of terms in the summation maybe finite or infinite. In the case where the sum has an infinite number of terms,the sum has meaning only if it converges. A common example of such a functionis the exponential of a matrix, which is defined byexp A =∞∑n=0A nn! . (8.38)This definition can, in turn, be used to define other functions such as sin A <strong>and</strong>cos A.8.6 The transpose of a matrixWe have seen that the components of a linear operator in a given coordinate systemcan be written in the <strong>for</strong>m of a matrix A. We will also find it useful, however,to consider the different (but clearly related) matrix <strong>for</strong>med by interchanging therows <strong>and</strong> columns of A. The matrix is called the transpose of A <strong>and</strong> is denotedby A T .255


MATRICES AND VECTOR SPACES◮Find the transpose of the matrix( )3 1 2A =.0 4 1By interchanging the rows <strong>and</strong> columns of A we immediately obtain⎛A T = ⎝ 3 0⎞1 4 ⎠ . ◭2 1It is obvious that if A is an M × N matrix then its transpose A T is a N × Mmatrix. As mentioned in section 8.3, the transpose of a column matrix is arow matrix <strong>and</strong> vice versa. An important use of column <strong>and</strong> row matrices isin the representation of the inner product of two real vectors in terms of theircomponents in a given basis. This notion is discussed fully in the next section,where it is extended to complex vectors.The transpose of the product of two matrices, (AB) T , is given by the productof their transposes taken in the reverse order, i.e.This is proved as follows:(AB) T = B T A T . (8.39)(AB) T ij =(AB) ji = ∑ k= ∑ kA jk B ki(A T ) kj (B T ) ik = ∑ k(B T ) ik (A T ) kj =(B T A T ) ij ,<strong>and</strong> the proof can be extended to the product of several matrices to give(ABC ···G) T = G T ···C T B T A T .8.7 The complex <strong>and</strong> Hermitian conjugates of a matrixTwo further matrices that can be derived from a given general M × N matrixare the complex conjugate, denoted by A ∗ ,<strong>and</strong>theHermitian conjugate, denotedby A † .The complex conjugate of a matrix A is the matrix obtained by taking thecomplex conjugate of each of the elements of A, i.e.(A ∗ ) ij =(A ij ) ∗ .Obviously if a matrix is real (i.e. it contains only real elements) then A ∗ = A.256


8.7 THE COMPLEX AND HERMITIAN CONJUGATES OF A MATRIX◮Find the complex conjugate of the matrix(1 2 3iA =1+i 1 0).By taking the complex conjugate of each element we obtain immediately( )1 2 −3iA ∗ =. ◭1 − i 1 0The Hermitian conjugate, or adjoint, of a matrix A is the transpose of itscomplex conjugate, or equivalently, the complex conjugate of its transpose, i.e.A † =(A ∗ ) T =(A T ) ∗ .We note that if A is real (<strong>and</strong> so A ∗ = A) thenA † = A T , <strong>and</strong> taking the Hermitianconjugate is equivalent to taking the transpose. Following the previous line ofargument <strong>for</strong> the transpose of the product of several matrices, the Hermitianconjugate of such a product can be shown to be given by(AB ···G) † = G † ···B † A † . (8.40)◮Find the Hermitian conjugate of the matrix(1 2 3iA =1+i 1 0).Taking the complex conjugate of A <strong>and</strong> then <strong>for</strong>ming the transpose we find⎛A † = ⎝ 1 1− i⎞2 1 ⎠ .−3i 0We obtain the same result, of course, if we first take the transpose of A <strong>and</strong>thentakethecomplex conjugate. ◭An important use of the Hermitian conjugate (or transpose in the real case)is in connection with the inner product of two vectors. Suppose that in a givenorthonormal basis the vectors a <strong>and</strong> b may be represented by the column matrices⎛a = ⎜⎝a 1a 2.a N⎞⎟⎠⎛<strong>and</strong> b = ⎜⎝b 1b 2.b N⎞⎟⎠ . (8.41)Taking the Hermitian conjugate of a, to give a row matrix, <strong>and</strong> multiplying (on257


MATRICES AND VECTOR SPACESthe right) by b we obtain⎛a † b =(a ∗ 1 a ∗ 2 ··· a ∗ N) ⎜⎝b 1b 2.b N⎞N⎟⎠ = ∑a ∗ i b i , (8.42)which is the expression <strong>for</strong> the inner product 〈a|b〉 in that basis. We note that <strong>for</strong>real vectors (8.42) reduces to a T b = ∑ Ni=1 a ib i .If the basis e i is not orthonormal, so that, in general,〈e i |e j 〉 = G ij ≠ δ ij ,then, from (8.18), the scalar product of a <strong>and</strong> b in terms of their components withrespect to this basis is given by〈a|b〉 =N∑i=1 j=1i=1N∑a ∗ i G ij b j = a † Gb,where G is the N × N matrix with elements G ij .8.8 The trace of a matrixFor a given matrix A, in the previous two sections we have considered variousother matrices that can be derived from it. However, sometimes one wishes toderive a single number from a matrix. The simplest example is the trace (or spur)of a square matrix, which is denoted by Tr A. This quantity is defined as the sumof the diagonal elements of the matrix,Tr A = A 11 + A 22 + ···+ A NN =N∑A ii . (8.43)It is clear that taking the trace is a linear operation so that, <strong>for</strong> example,Tr(A ± B) =TrA ± Tr B.A very useful property of traces is that the trace of the product of two matricesis independent of the order of their multiplication; this results holds whether ornot the matrices commute <strong>and</strong> is proved as follows:N∑N∑ N∑N∑ N∑N∑Tr AB = (AB) ii = A ij B ji = B ji A ij =i=1i=1 j=1i=1 j=1i=1j=1(BA) jj =TrBA.(8.44)The result can be extended to the product of several matrices. For example, from(8.44), we immediately findTr ABC =TrBCA =TrCAB,258


8.9 THE DETERMINANT OF A MATRIXwhich shows that the trace of a multiple product is invariant under cyclicpermutations of the matrices in the product. Other easily derived properties ofthe trace are, <strong>for</strong> example, Tr A T =TrA <strong>and</strong> Tr A † =(TrA) ∗ .8.9 The determinant of a matrixFor a given matrix A, the determinant det A (like the trace) is a single number (oralgebraic expression) that depends upon the elements of A. Also like the trace,the determinant is defined only <strong>for</strong> square matrices. If, <strong>for</strong> example, A is a 3 × 3matrix then its determinant, of order 3, is denoted bydet A = |A| =∣A 11 A 12 A 13A 21 A 22 A 23A 31 A 32 A 33∣ ∣∣∣∣∣. (8.45)In order to calculate the value of a determinant, we first need to introducethe notions of the minor <strong>and</strong> the cofactor of an element of a matrix. (Weshall see that we can use the cofactors to write an order-3 determinant as theweighted sum of three order-2 determinants, thereby simplifying its evaluation.)The minor M ij of the element A ij of an N × N matrix A is the determinant ofthe (N − 1) × (N − 1) matrix obtained by removing all the elements of the ithrow <strong>and</strong> jth column of A; the associated cofactor, C ij , is found by multiplyingthe minor by (−1) i+j .◮Find the cofactor of the element A 23 of the matrix⎛A = ⎝ A ⎞11 A 12 A 13A 21 A 22 A 23⎠ .A 31 A 32 A 33Removing all the elements of the second row <strong>and</strong> third column of A <strong>and</strong> <strong>for</strong>ming thedeterminant of the remaining terms gives the minorM 23 =∣ A ∣11 A 12 ∣∣∣.A 31 A 32Multiplying the minor by (−1) 2+3 =(−1) 5 = −1 givesC 23 = −∣ A ∣11 A 12 ∣∣∣. ◭A 31 A 32We now define a determinant as the sum of the products of the elements of anyrow or column <strong>and</strong> their corresponding cofactors, e.g.A 21 C 21 + A 22 C 22 + A 23 C 23 orA 13 C 13 + A 23 C 23 + A 33 C 33 . Such a sum is called a Laplace expansion. For example,in the first of these expansions, using the elements of the second row of the259


MATRICES AND VECTOR SPACESdeterminant defined by (8.45) <strong>and</strong> their corresponding cofactors, we write |A| asthe Laplace expansion|A| = A 21 (−1) (2+1) M 21 + A 22 (−1) (2+2) M 22 + A 23 (−1) (2+3) M 23= −A 21∣ ∣∣∣ A 12 A 13A 32 A 33∣ ∣∣∣+ A 22∣ ∣∣∣ A 11 A 13A 31 A 33∣ ∣∣∣− A 23∣ ∣∣∣ A 11 A 12A 31 A 32∣ ∣∣∣.We will see later that the value of the determinant is independent of the rowor column chosen. Of course, we have not yet determined the value of |A| but,rather, written it as the weighted sum of three determinants of order 2. However,applying again the definition of a determinant, we can evaluate each of theorder-2 determinants.◮Evaluate the determinant ∣ ∣∣∣ A 12 A 13A 32 A 33∣ ∣∣∣.By considering the products of the elements of the first row in the determinant, <strong>and</strong> theircorresponding cofactors, we find∣ A ∣12 A 13 ∣∣∣= AA 32 A 12 (−1) (1+1) |A 33 | + A 13 (−1) (1+2) |A 32 |33= A 12 A 33 − A 13 A 32 ,where the values of the order-1 determinants |A 33 | <strong>and</strong> |A 32 | are defined to be A 33 <strong>and</strong> A 32respectively. It must be remembered that the determinant is not the same as the modulus,e.g. det (−2) = |−2| = −2, not 2. ◭We can now combine all the above results to show that the value of thedeterminant (8.45) is given by|A| = −A 21 (A 12 A 33 − A 13 A 32 )+A 22 (A 11 A 33 − A 13 A 31 )− A 23 (A 11 A 32 − A 12 A 31 ) (8.46)= A 11 (A 22 A 33 − A 23 A 32 )+A 12 (A 23 A 31 − A 21 A 33 )+ A 13 (A 21 A 32 − A 22 A 31 ), (8.47)where the final expression gives the <strong>for</strong>m in which the determinant is usuallyremembered <strong>and</strong> is the <strong>for</strong>m that is obtained immediately by considering theLaplace expansion using the first row of the determinant. The last equality, whichessentially rearranges a Laplace expansion using the second row into one usingthe first row, supports our assertion that the value of the determinant is unaffectedby which row or column is chosen <strong>for</strong> the expansion.260


8.9 THE DETERMINANT OF A MATRIX◮Suppose the rows of a real 3 × 3 matrix A are interpreted as the components in a givenbasis of three (three-component) vectors a, b <strong>and</strong> c. Show that one can write the determinantof A as|A| = a · (b × c).If one writes the rows of A as the components in a given basis of three vectors a, b <strong>and</strong> c,we have from (8.47) that∣ a 1 a 2 a 3 ∣∣∣∣∣|A| =b 1 b 2 b 3 = a 1 (b 2 c 3 − b 3 c 2 )+a 2 (b 3 c 1 − b 1 c 3 )+a 3 (b 1 c 2 − b 2 c 1 ).∣ c 1 c 2 c 3From expression (7.34) <strong>for</strong> the scalar triple product given in subsection 7.6.3, it followsthat we may write the determinant as|A| = a · (b × c). (8.48)In other words, |A| is the volume of the parallelepiped defined by the vectors a, b <strong>and</strong>c. (One could equally well interpret the columns of the matrix A as the components ofthree vectors, <strong>and</strong> result (8.48) would still hold.) This result provides a more memorable(<strong>and</strong> more meaningful) expression than (8.47) <strong>for</strong> the value of a 3 × 3 determinant. Indeed,using this geometrical interpretation, we see immediately that, if the vectors a 1 , a 2 , a 3 arenot linearly independent then the value of the determinant vanishes: |A| =0.◭The evaluation of determinants of order greater than 3 follows the same generalmethod as that presented above, in that it relies on successively reducing the orderof the determinant by writing it as a Laplace expansion. Thus, a determinantof order 4 is first written as a sum of four determinants of order 3, whichare then evaluated using the above method. For higher-order determinants, onecannot write down directly a simple geometrical expression <strong>for</strong> |A| analogous tothat given in (8.48). Nevertheless, it is still true that if the rows or columns ofthe N × N matrix A are interpreted as the components in a given basis of N(N-component) vectors a 1 , a 2 ,...,a N , then the determinant |A| vanishes if thesevectors are not all linearly independent.8.9.1 Properties of determinantsA number of properties of determinants follow straight<strong>for</strong>wardly from the definitionof det A; their use will often reduce the labour of evaluating a determinant.We present them here without specific proofs, though they all follow readily fromthe alternative <strong>for</strong>m <strong>for</strong> a determinant, given in equation (26.29) on page 942,<strong>and</strong> expressed in terms of the Levi–Civita symbol ɛ ijk (see exercise 26.9).(i) Determinant of the transpose. The transpose matrix A T (which, we recall,is obtained by interchanging the rows <strong>and</strong> columns of A) has the samedeterminant as A itself, i.e.|A T | = |A|. (8.49)261


MATRICES AND VECTOR SPACESIt follows that any theorem established <strong>for</strong> the rows of A will apply to thecolumns as well, <strong>and</strong> vice versa.(ii) Determinant of the complex <strong>and</strong> Hermitian conjugate. It is clear that thematrix A ∗ obtained by taking the complex conjugate of each element of Ahas the determinant |A ∗ | = |A| ∗ . Combining this result with (8.49), we findthat|A † | = |(A ∗ ) T | = |A ∗ | = |A| ∗ . (8.50)(iii) Interchanging two rows or two columns. If two rows (columns) of A areinterchanged, its determinant changes sign but is unaltered in magnitude.(iv) Removing factors. If all the elements of a single row (column) of A havea common factor, λ, then this factor may be removed; the value of thedeterminant is given by the product of the remaining determinant <strong>and</strong> λ.Clearly this implies that if all the elements of any row (column) are zerothen |A| = 0. It also follows that if every element of the N × N matrix Ais multiplied by a constant factor λ then|λA| = λ N |A|. (8.51)(v) Identical rows or columns. If any two rows (columns) of A are identical orare multiples of one another, then it can be shown that |A| =0.(vi) Adding a constant multiple of one row (column) to another. The determinantof a matrix is unchanged in value by adding to the elements of one row(column) any fixed multiple of the elements of another row (column).(vii) Determinant of a product. IfA <strong>and</strong> B are square matrices of the same orderthen|AB| = |A||B| = |BA|. (8.52)A simple extension of this property gives, <strong>for</strong> example,|AB ···G| = |A||B| ···|G| = |A||G| ···|B| = |A ···GB|,which shows that the determinant is invariant under permutation of thematrices in a multiple product.There is no explicit procedure <strong>for</strong> using the above results in the evaluation ofany given determinant, <strong>and</strong> judging the quickest route to an answer is a matterof experience. A general guide is to try to reduce all terms but one in a row orcolumn to zero <strong>and</strong> hence in effect to obtain a determinant of smaller size. Thesteps taken in evaluating the determinant in the example below are certainly notthe fastest, but they have been chosen in order to illustrate the use of most of theproperties listed above.262


8.10 THE INVERSE OF A MATRIX◮Evaluate the determinant|A| =∣1 0 2 30 1 −2 13 −3 4 −2−2 1 −2 −1.∣Taking a factor 2 out of the third column <strong>and</strong> then adding the second column to the thirdgives1 0 1 31 0 1 30 1 −1 10 1 0 1|A| =23 −3 2 −2=23 −3 −1 −2.∣ −2 1 −1 −1 ∣ ∣ −2 1 0 −1 ∣Subtracting the second column from the fourth gives1 0 1 30 1 0 0|A| =23 −3 −1 1.∣ −2 1 0 −2 ∣We now note that the second row has only one non-zero element <strong>and</strong> so the determinantmay conveniently be written as a Laplace expansion, i.e.∣ ∣ ∣∣∣∣∣ 1 1 3∣∣∣∣∣ 4 0 4|A| =2× 1 × (−1) 2+2 3 −1 1−2 0 −2 ∣ =2 3 −1 1−2 0 −2 ∣ ,where the last equality follows by adding the second row to the first. It can now be seenthat the first row is minus twice the third, <strong>and</strong> so the value of the determinant is zero, byproperty (v) above. ◭8.10 The inverse of a matrixOur first use of determinants will be in defining the inverse of a matrix. If wewere dealing with ordinary numbers we would consider the relation P = AB asequivalent to B = P/A, provided that A ≠ 0. However, if A, B <strong>and</strong> P are matricesthen this notation does not have an obvious meaning. What we really want toknow is whether an explicit <strong>for</strong>mula <strong>for</strong> B can be obtained in terms of A <strong>and</strong>P. It will be shown that this is possible <strong>for</strong> those cases in which |A| ≠0.Asquare matrix whose determinant is zero is called a singular matrix; otherwise itis non-singular. We will show that if A is non-singular we can define a matrix,denoted by A −1 <strong>and</strong> called the inverse of A, which has the property that if AB = Pthen B = A −1 P.Inwords,B can be obtained by multiplying P from the left byA −1 . Analogously, if B is non-singular then, by multiplication from the right,A = PB −1 .It is clear thatAI = A ⇒ I = A −1 A, (8.53)where I is the unit matrix, <strong>and</strong> so A −1 A = I = AA −1 . These statements are263


MATRICES AND VECTOR SPACESequivalent to saying that if we first multiply a matrix, B say, by A <strong>and</strong> thenmultiply by the inverse A −1 , we end up with the matrix we started with, i.e.A −1 AB = B. (8.54)This justifies our use of the term inverse. It is also clear that the inverse is onlydefined <strong>for</strong> square matrices.So far we have only defined what we mean by the inverse of a matrix. Actuallyfinding the inverse of a matrix A may be carried out in a number of ways. We willshow that one method is to construct first the matrix C containing the cofactorsof the elements of A, as discussed in the last subsection. Then the required inverseA −1 can be found by <strong>for</strong>ming the transpose of C <strong>and</strong> dividing by the determinantof A. Thus the elements of the inverse A −1 are given by(A −1 ) ik = (C)T ik|A|= C ki|A| . (8.55)That this procedure does indeed result in the inverse may be seen by consideringthe components of A −1 A,i.e.(A −1 A) ij = ∑ k(A −1 ) ik (A) kj = ∑ kC ki|A| A kj = |A||A| δ ij. (8.56)The last equality in (8.56) relies on the property∑C ki A kj = |A|δ ij ; (8.57)kthis can be proved by considering the matrix A ′ obtained from the original matrixA when the ith column of A is replaced by one of the other columns, say the jth.Thus A ′ is a matrix with two identical columns <strong>and</strong> so has zero determinant.However, replacing the ith column by another does not change the cofactors C kiof the elements in the ith column, which are there<strong>for</strong>e the same in A <strong>and</strong> A ′ .Recalling the Laplace expansion of a determinant, i.e.|A| = ∑ kA ki C ki ,we obtain0=|A ′ | = ∑ kA ′ kiC ′ ki = ∑ kA kj C ki , i ≠ j,which together with the Laplace expansion itself may be summarised by (8.57).It is immediately obvious from (8.55) that the inverse of a matrix is not definedif the matrix is singular (i.e. if |A| =0).264


8.10 THE INVERSE OF A MATRIX◮Find the inverse of the matrix⎛A = 1 −2 −2⎝ 2 4 3⎞⎠ .−3 3 2We first determine |A|:|A| =2[−2(2) − (−2)3] + 4[(−2)(−3) − (1)(2)] + 3[(1)(3) − (−2)(−3)]=11. (8.58)This is non-zero <strong>and</strong> so an inverse matrix can be constructed. To do this we need thematrix of the cofactors, C, <strong>and</strong> hence C T . We find⎛C = ⎝ 2 4 −3⎞⎛1 13 −18 ⎠ <strong>and</strong> C T = ⎝ 2 1 −2⎞4 13 7 ⎠ ,−2 7 −8−3 −18 −8<strong>and</strong> henceA −1 = CT|A| = 111⎛⎝ 4 13 72 1 −2−3 −18 −8⎞⎠ . ◭ (8.59)For a 2 × 2 matrix, the inverse has a particularly simple <strong>for</strong>m. If the matrix is( )A11 AA =12A 21 A 22then its determinant |A| is given by |A| = A 11 A 22 − A 12 A 21 ,<strong>and</strong>thematrixofcofactors is( )A22 −AC =21.−A 12 A 11Thus the inverse of A is given by( )A −1 = CT|A| = 1A22 −A 12. (8.60)A 11 A 22 − A 12 A 21 −A 21 A 11It can be seen that the transposed matrix of cofactors <strong>for</strong> a 2 × 2matrixisthesame as the matrix <strong>for</strong>med by swapping the elements on the leading diagonal(A 11 <strong>and</strong> A 22 ) <strong>and</strong> changing the signs of the other two elements (A 12 <strong>and</strong> A 21 ).This is completely general <strong>for</strong> a 2 × 2 matrix <strong>and</strong> is easy to remember.The following are some further useful properties related to the inverse matrix265


MATRICES AND VECTOR SPACES<strong>and</strong> may be straight<strong>for</strong>wardly derived.(i) (A −1 ) −1 = A.(ii) (A T ) −1 =(A −1 ) T .(iii) (A † ) −1 =(A −1 ) † .(iv) (AB) −1 = B −1 A −1 .(v) (AB ···G) −1 = G −1 ···B −1 A −1 .◮Prove the properties (i)–(v) stated above.We begin by writing down the fundamental expression defining the inverse of a nonsingularsquare matrix A:AA −1 = I = A −1 A. (8.61)Property (i). This follows immediately from the expression (8.61).Property (ii). Taking the transpose of each expression in (8.61) gives(AA −1 ) T = I T =(A −1 A) T .Using the result (8.39) <strong>for</strong> the transpose of a product of matrices <strong>and</strong> noting that I T = I,we find(A −1 ) T A T = I = A T (A −1 ) T .However, from (8.61), this implies (A −1 ) T =(A T ) −1 <strong>and</strong> hence proves result (ii) above.Property (iii). This may be proved in an analogous way to property (ii), by replacing thetransposes in (ii) by Hermitian conjugates <strong>and</strong> using the result (8.40) <strong>for</strong> the Hermitianconjugate of a product of matrices.Property (iv). Using (8.61), we may write(AB)(AB) −1 = I = (AB) −1 (AB),From the left-h<strong>and</strong> equality it follows, by multiplying on the left by A −1 ,thatA −1 AB(AB) −1 = A −1 I <strong>and</strong> hence B(AB) −1 = A −1 .Now multiplying on the left by B −1 givesB −1 B(AB) −1 = B −1 A −1 ,<strong>and</strong> hence the stated result.Property (v). Finally, result (iv) may extended to case (v) in a straight<strong>for</strong>ward manner.For example, using result (iv) twice we find(ABC) −1 = (BC) −1 A −1 = C −1 B −1 A −1 . ◭We conclude this section by noting that the determinant |A −1 | of the inversematrix can be expressed very simply in terms of the determinant |A| of the matrixitself. Again we start with the fundamental expression (8.61). Then, using theproperty (8.52) <strong>for</strong> the determinant of a product, we find|AA −1 | = |A||A −1 | = |I|.It is straight<strong>for</strong>ward to show by Laplace expansion that |I| = 1, <strong>and</strong> so we arriveat the useful result|A −1 | = 1|A| . (8.62)266


8.11 THE RANK OF A MATRIX8.11 The rank of a matrixThe rank of a general M × N matrix is an important concept, particularly inthe solution of sets of simultaneous linear equations, to be discussed in the nextsection, <strong>and</strong> we now discuss it in some detail. Like the trace <strong>and</strong> determinant,the rank of matrix A is a single number (or algebraic expression) that dependson the elements of A. Unlike the trace <strong>and</strong> determinant, however, the rank of amatrix can be defined even when A is not square. As we shall see, there are twoequivalent definitions of the rank of a general matrix.Firstly, the rank of a matrix may be defined in terms of the linear independenceof vectors. Suppose that the columns of an M × N matrix are interpreted asthe components in a given basis of N (M-component) vectors v 1 , v 2 ,...,v N ,asfollows:⎛A = ⎝↑ ↑ ↑v 1 v 2 ... v N↓ ↓ ↓Then the rank of A, denoted by rank A or by R(A), is defined as the numberof linearly independent vectors in the set v 1 , v 2 ,...,v N , <strong>and</strong> equals the dimensionof the vector space spanned by those vectors. Alternatively, we may consider therows of A to contain the components in a given basis of the M (N-component)vectors w 1 , w 2 ,...,w M as follows:⎛A = ⎜⎝← w 1 →← w 2 →.← w M →It may then be shown § that the rank of A is also equal to the number oflinearly independent vectors in the set w 1 , w 2 ,...,w M . From this definition it isshould be clear that the rank of A is unaffected by the exchange of two rows(or two columns) or by the multiplication of a row (or column) by a constant.Furthermore, suppose that a constant multiple of one row (column) is added toanother row (column): <strong>for</strong> example, we might replace the row w i by w i + cw j .This also has no effect on the number of linearly independent rows <strong>and</strong> so leavesthe rank of A unchanged. We may use these properties to evaluate the rank of agiven matrix.A second (equivalent) definition of the rank of a matrix may be given <strong>and</strong> usesthe concept of submatrices. A submatrix of A is any matrix that can be <strong>for</strong>medfrom the elements of A by ignoring one, or more than one, row or column. It⎞⎠ .⎞⎟⎠ .§ For a fuller discussion, see, <strong>for</strong> example, C. D. Cantrell, Modern <strong>Mathematical</strong> <strong>Methods</strong> <strong>for</strong> Physicists<strong>and</strong> Engineers (Cambridge: Cambridge University Press, 2000), chapter 6.267


MATRICES AND VECTOR SPACESmay be shown that the rank of a general M × N matrix is equal to the size ofthe largest square submatrix of A whose determinant is non-zero. There<strong>for</strong>e, if amatrix A has an r × r submatrix S with |S| ≠ 0, but no (r +1)× (r + 1) submatrixwith non-zero determinant then the rank of the matrix is r. From either definitionit is clear that the rank of A is less than or equal to the smaller of M <strong>and</strong> N.◮Determine the rank of the matrix⎛A = ⎝ 1 1 0 −22 0 2 24 1 3 1⎞⎠ .The largest possible square submatrices of A must be of dimension 3 × 3. Clearly, Apossesses four such submatrices, the determinants of which are given by∣ 1 1 0∣∣∣∣∣ 1 1 −22 0 2∣ 4 1 3 ∣ =0, 2 0 24 1 1 ∣ =0,∣1 0 −22 2 24 3 1∣ ∣∣∣∣∣ 1 0 −2∣ =0, 0 2 21 3 1∣ =0.(In each case the determinant may be evaluated as described in subsection 8.9.1.)The next largest square submatrices of A are of dimension 2 × 2. Consider, <strong>for</strong> example,the 2 × 2 submatrix <strong>for</strong>med by ignoring the third row <strong>and</strong> the third <strong>and</strong> fourth columnsof A; this has determinant∣ 1 12 0 ∣ =1× 0 − 2 × 1=−2.Since its determinant is non-zero, A is of rank 2 <strong>and</strong> we need not consider any other 2 × 2submatrix. ◭In the special case in which the matrix A is a square N×N matrix, by comparingeither of the above definitions of rank with our discussion of determinants insection 8.9, we see that |A| = 0 unless the rank of A is N. Inotherwords,A issingular unless R(A) =N.8.12 Special types of square matrixMatrices that are square, i.e. N × N, are very common in physical applications.We now consider some special <strong>for</strong>ms of square matrix that are of particularimportance.8.12.1 Diagonal matricesThe unit matrix, which we have already encountered, is an example of a diagonalmatrix. Such matrices are characterised by having non-zero elements only on the268


8.12 SPECIAL TYPES OF SQUARE MATRIXleading diagonal, i.e. only elements A ij with i = j may be non-zero. For example,⎛1 0 0⎞A = ⎝ 0 2 0 ⎠ ,0 0 −3is a 3 × 3 diagonal matrix. Such a matrix is often denoted by A = diag (1, 2, −3).By per<strong>for</strong>ming a Laplace expansion, it is easily shown that the determinant of anN × N diagonal matrix is equal to the product of the diagonal elements. Thus, ifthe matrix has the <strong>for</strong>m A = diag(A 11 ,A 22 ,...,A NN )then|A| = A 11 A 22 ···A NN . (8.63)Moreover, it is also straight<strong>for</strong>ward to show that the inverse of A is also adiagonal matrix given by( )1A −1 1 1= diag , ,..., .A 11 A 22 A NNFinally, we note that, if two matrices A <strong>and</strong> B are both diagonal then they havethe useful property that their product is commutative:This is not true <strong>for</strong> matrices in general.AB = BA.8.12.2 Lower <strong>and</strong> upper triangular matricesA square matrix A is called lower triangular if all the elements above the principaldiagonal are zero. For example, the general <strong>for</strong>m <strong>for</strong> a 3 × 3 lower triangularmatrix is⎛A 11 0 0⎞A 31 A 32 A 33A = ⎝ A 21 A 22 0 ⎠ ,where the elements A ij may be zero or non-zero. Similarly an upper triangularsquare matrix is one <strong>for</strong> which all the elements below the principal diagonal arezero. The general 3 × 3 <strong>for</strong>m is thus⎛A 11 A 12 A 13⎞0 0 A 33A = ⎝ 0 A 22 A 23⎠ .By per<strong>for</strong>ming a Laplace expansion, it is straight<strong>for</strong>ward to show that, in thegeneral N × N case, the determinant of an upper or lower triangular matrix isequal to the product of its diagonal elements,|A| = A 11 A 22 ···A NN . (8.64)269


MATRICES AND VECTOR SPACESClearly result (8.63) <strong>for</strong> diagonal matrices is a special case of this result. Moreover,it may be shown that the inverse of a non-singular lower (upper) triangular matrixis also lower (upper) triangular.8.12.3 Symmetric <strong>and</strong> antisymmetric matricesA square matrix A of order N with the property A = A T is said to be symmetric.Similarly a matrix <strong>for</strong> which A = −A T is said to be anti- or skew-symmetric<strong>and</strong> its diagonal elements a 11 ,a 22 ,...,a NN are necessarily zero. Moreover, if A is(anti-)symmetric then so too is its inverse A −1 . This is easily proved by notingthat if A = ±A T then(A −1 ) T =(A T ) −1 = ±A −1 .Any N × N matrix A can be written as the sum of a symmetric <strong>and</strong> anantisymmetric matrix, since we may writeA = 1 2 (A + AT )+ 1 2 (A − AT )=B + C,where clearly B = B T <strong>and</strong> C = −C T . The matrix B is there<strong>for</strong>e called thesymmetric part of A, <strong>and</strong>C is the antisymmetric part.◮If A is an N × N antisymmetric matrix, show that |A| =0if N is odd.If A is antisymmetric then A T = −A. Using the properties of determinants (8.49) <strong>and</strong>(8.51), we have|A| = |A T | = |−A| =(−1) N |A|.Thus, if N is odd then |A| = −|A|, which implies that |A| =0.◭8.12.4 Orthogonal matricesA non-singular matrix with the property that its transpose is also its inverse,A T = A −1 , (8.65)is called an orthogonal matrix. It follows immediately that the inverse of anorthogonal matrix is also orthogonal, since(A −1 ) T =(A T ) −1 =(A −1 ) −1 .Moreover, since <strong>for</strong> an orthogonal matrix A T A = I, we have|A T A| = |A T ||A| = |A| 2 = |I| =1.Thus the determinant of an orthogonal matrix must be |A| = ±1.An orthogonal matrix represents, in a particular basis, a linear operator thatleaves the norms (lengths) of real vectors unchanged, as we will now show.270


8.12 SPECIAL TYPES OF SQUARE MATRIXSuppose that y = A x is represented in some coordinate system by the matrixequation y = Ax; then〈y|y〉 is given in this coordinate system byy T y = x T A T Ax = x T x.Hence 〈y|y〉 = 〈x|x〉, showing that the action of a linear operator represented byan orthogonal matrix does not change the norm of a real vector.8.12.5 Hermitian <strong>and</strong> anti-Hermitian matricesAn Hermitian matrix is one that satisfies A = A † ,whereA † is the Hermitian conjugatediscussed in section 8.7. Similarly if A † = −A, thenA is called anti-Hermitian.A real (anti-)symmetric matrix is a special case of an (anti-)Hermitian matrix, inwhich all the elements of the matrix are real. Also, if A is an (anti-)Hermitianmatrix then so too is its inverse A −1 ,since(A −1 ) † =(A † ) −1 = ±A −1 .Any N × N matrix A can be written as the sum of an Hermitian matrix <strong>and</strong>an anti-Hermitian matrix, sinceA = 1 2 (A + A† )+ 1 2 (A − A† )=B + C,where clearly B = B † <strong>and</strong> C = −C † . The matrix B is called the Hermitian part ofA, <strong>and</strong>C is called the anti-Hermitian part.8.12.6 Unitary matricesA unitary matrix A is defined as one <strong>for</strong> whichA † = A −1 . (8.66)Clearly, if A is real then A † = A T , showing that a real orthogonal matrix is aspecial case of a unitary matrix, one in which all the elements are real. We notethat the inverse A −1 of a unitary is also unitary, since(A −1 ) † =(A † ) −1 =(A −1 ) −1 .Moreover, since <strong>for</strong> a unitary matrix A † A = I, we have|A † A| = |A † ||A| = |A| ∗ |A| = |I| =1.Thus the determinant of a unitary matrix has unit modulus.A unitary matrix represents, in a particular basis, a linear operator that leavesthe norms (lengths) of complex vectors unchanged. If y = A x is represented insome coordinate system by the matrix equation y = Ax then 〈y|y〉 is given in thiscoordinate system byy † y = x † A † Ax = x † x.271


MATRICES AND VECTOR SPACESHence 〈y|y〉 = 〈x|x〉, showing that the action of the linear operator represented bya unitary matrix does not change the norm of a complex vector. The action of aunitary matrix on a complex column matrix thus parallels that of an orthogonalmatrix acting on a real column matrix.8.12.7 Normal matricesA final important set of special matrices consists of the normal matrices, <strong>for</strong> whichAA † = A † A,i.e. a normal matrix is one that commutes with its Hermitian conjugate.We can easily show that Hermitian matrices <strong>and</strong> unitary matrices (or symmetricmatrices <strong>and</strong> orthogonal matrices in the real case) are examples of normalmatrices. For an Hermitian matrix, A = A † <strong>and</strong> soAA † = AA = A † A.Similarly, <strong>for</strong> a unitary matrix, A −1 = A † <strong>and</strong> soAA † = AA −1 = A −1 A = A † A.Finally, we note that, if A is normal then so too is its inverse A −1 ,sinceA −1 (A −1 ) † = A −1 (A † ) −1 =(A † A) −1 =(AA † ) −1 =(A † ) −1 A −1 =(A −1 ) † A −1 .This broad class of matrices is important in the discussion of eigenvectors <strong>and</strong>eigenvalues in the next section.8.13 Eigenvectors <strong>and</strong> eigenvaluesSuppose that a linear operator A trans<strong>for</strong>ms vectors x in an N-dimensionalvector space into other vectors A x in the same space. The possibility then arisesthat there exist vectors x each of which is trans<strong>for</strong>med by A into a multiple ofitself. Such vectors would have to satisfyA x = λx. (8.67)Any non-zero vector x that satisfies (8.67) <strong>for</strong> some value of λ is called aneigenvector of the linear operator A ,<strong>and</strong>λ is called the corresponding eigenvalue.As will be discussed below, in general the operator A has N independenteigenvectors x i , with eigenvalues λ i .Theλ i are not necessarily all distinct.If we choose a particular basis in the vector space, we can write (8.67) in termsof the components of A <strong>and</strong> x with respect to this basis as the matrix equationAx = λx, (8.68)where A is an N × N matrix. The column matrices x that satisfy (8.68) obviously272


8.13 EIGENVECTORS AND EIGENVALUESrepresent the eigenvectors x of A in our chosen coordinate system. Conventionally,these column matrices are also referred to as the eigenvectors of the matrixA. § Clearly, if x is an eigenvector of A (with some eigenvalue λ) then any scalarmultiple µx is also an eigenvector with the same eigenvalue. We there<strong>for</strong>e oftenuse normalised eigenvectors, <strong>for</strong> whichx † x =1(note that x † x corresponds to the inner product 〈x|x〉 in our basis). Any eigenvectorx can be normalised by dividing all its components by the scalar (x † x) 1/2 .As will be seen, the problem of finding the eigenvalues <strong>and</strong> correspondingeigenvectors of a square matrix A plays an important role in many physicalinvestigations. Throughout this chapter we denote the ith eigenvector of a squarematrix A by x i <strong>and</strong> the corresponding eigenvalue by λ i . This superscript notation<strong>for</strong> eigenvectors is used to avoid any confusion with components.◮A non-singular matrix A has eigenvalues λ i <strong>and</strong> eigenvectors x i . Find the eigenvalues <strong>and</strong>eigenvectors of the inverse matrix A −1 .The eigenvalues <strong>and</strong> eigenvectors of A satisfyAx i = λ i x i .Left-multiplying both sides of this equation by A −1 , we findA −1 Ax i = λ i A −1 x i .Since A −1 A = I, on rearranging we obtainA −1 x i = 1 λ ix i .Thus, we see that A −1 has the same eigenvectors x i as does A, but the correspondingeigenvalues are 1/λ i . ◭In the remainder of this section we will discuss some useful results concerningthe eigenvectors <strong>and</strong> eigenvalues of certain special (though commonly occurring)square matrices. The results will be established <strong>for</strong> matrices whose elements maybe complex; the corresponding properties <strong>for</strong> real matrices may be obtained asspecial cases.8.13.1 Eigenvectors <strong>and</strong> eigenvalues of a normal matrixIn subsection 8.12.7 we defined a normal matrix A as one that commutes with itsHermitian conjugate, so thatA † A = AA † .§ In this context, when referring to linear combinations of eigenvectors x we will normally use theterm ‘vector’.273


MATRICES AND VECTOR SPACESWe also showed that both Hermitian <strong>and</strong> unitary matrices (or symmetric <strong>and</strong>orthogonal matrices in the real case) are examples of normal matrices. We nowdiscuss the properties of the eigenvectors <strong>and</strong> eigenvalues of a normal matrix.If x is an eigenvector of a normal matrix A with corresponding eigenvalue λthen Ax = λx, or equivalently,(A − λI)x = 0. (8.69)Denoting B = A−λI, (8.69) becomes Bx = 0 <strong>and</strong>, taking the Hermitian conjugate,we also haveFrom (8.69) <strong>and</strong> (8.70) we then haveHowever, the product B † B is given by(Bx) † = x † B † = 0. (8.70)x † B † Bx = 0. (8.71)B † B =(A − λI) † (A − λI) =(A † − λ ∗ I)(A − λI) =A † A − λ ∗ A − λA † + λλ ∗ .Now since A is normal, AA † = A † A <strong>and</strong> soB † B = AA † − λ ∗ A − λA † + λλ ∗ =(A − λI)(A − λI) † = BB † ,<strong>and</strong> hence B is also normal. From (8.71) we then findx † B † Bx = x † BB † x =(B † x) † B † x = 0,from which we obtainB † x =(A † − λ ∗ I)x = 0.There<strong>for</strong>e, <strong>for</strong> a normal matrix A, the eigenvalues of A † are the complex conjugatesof the eigenvalues of A.Let us now consider two eigenvectors x i <strong>and</strong> x j of a normal matrix A correspondingto two different eigenvalues λ i <strong>and</strong> λ j . We then haveMultiplying (8.73) on the left by (x i ) † we obtainHowever, on the LHS of (8.74) we haveAx i = λ i x i , (8.72)Ax j = λ j x j . (8.73)(x i ) † Ax j = λ j (x i ) † x j . (8.74)(x i ) † A =(A † x i ) † =(λ ∗ i x i ) † = λ i (x i ) † , (8.75)where we have used (8.40) <strong>and</strong> the property just proved <strong>for</strong> a normal matrix to274


8.13 EIGENVECTORS AND EIGENVALUESwrite A † x i = λ ∗ i xi . From (8.74) <strong>and</strong> (8.75) we have(λ i − λ j )(x i ) † x j = 0. (8.76)Thus, if λ i ≠ λ j the eigenvectors x i <strong>and</strong> x j must be orthogonal, i.e.(x i ) † x j = 0.It follows immediately from (8.76) that if all N eigenvalues of a normal matrixA are distinct then all N eigenvectors of A are mutually orthogonal. If, however,two or more eigenvalues are the same then further consideration is required. Aneigenvalue corresponding to two or more different eigenvectors (i.e. they are notsimply multiples of one another) is said to be degenerate. Suppose that λ 1 is k-folddegenerate, i.e.Ax i = λ 1 x i <strong>for</strong> i =1, 2,...,k, (8.77)but that it is different from any of λ k+1 , λ k+2 , etc. Then any linear combinationof these x i is also an eigenvector with eigenvalue λ 1 ,since,<strong>for</strong>z = ∑ ki=1 c ix i ,Az ≡ Ak∑c i x i =i=1k∑c i Ax i =i=1k∑c i λ 1 x i = λ 1 z. (8.78)If the x i defined in (8.77) are not already mutually orthogonal then we canconstruct new eigenvectors z i that are orthogonal by the following procedure:i=1z 1 = x 1 ,[z 2 = x 2 − (ẑ 1 ) † x 2] ẑ 1 ,[z 3 = x 3 − (ẑ 2 ) † x 3] [ẑ 2 − (ẑ 1 ) † x 3] ẑ 1 ,.z k = x k −[(ẑ k−1 ) † x k] [ẑ k−1 − ···− (ẑ 1 ) † x k] ẑ 1 .In this procedure, known as Gram–Schmidt orthogonalisation, each new eigenvectorz i is normalised to give the unit vector ẑ i be<strong>for</strong>e proceeding to the constructionof the next one (the normalisation is carried out by dividing each element ofthe vector z i by [(z i ) † z i ] 1/2 ). Note that each factor in brackets (ẑ m ) † x n is a scalarproduct <strong>and</strong> thus only a number. It follows that, as shown in (8.78), each vectorz i so constructed is an eigenvector of A with eigenvalue λ 1 <strong>and</strong> will remain soon normalisation. It is straight<strong>for</strong>ward to check that, provided the previous neweigenvectors have been normalised as prescribed, each z i is orthogonal to all itspredecessors. (In practice, however, the method is laborious <strong>and</strong> the example insubsection 8.14.1 gives a less rigorous but considerably quicker way.)There<strong>for</strong>e, even if A has some degenerate eigenvalues we can by constructionobtain a set of N mutually orthogonal eigenvectors. Moreover, it may be shown(although the proof is beyond the scope of this book) that these eigenvectorsare complete in that they <strong>for</strong>m a basis <strong>for</strong> the N-dimensional vector space. As275


MATRICES AND VECTOR SPACESa result any arbitrary vector y can be expressed as a linear combination of theeigenvectors x i :y =N∑a i x i , (8.79)i=1where a i =(x i ) † y. Thus, the eigenvectors <strong>for</strong>m an orthogonal basis <strong>for</strong> the vectorspace. By normalising the eigenvectors so that (x i ) † x i = 1 this basis is madeorthonormal.◮Show that a normal matrix A can be written in terms of its eigenvalues λ i <strong>and</strong> orthonormaleigenvectors x i asN∑A = λ i x i (x i ) † . (8.80)i=1The key to proving the validity of (8.80) is to show that both sides of the expression givethesameresultwhenactingonanarbitaryvectory. SinceA is normal, we may exp<strong>and</strong> yin terms of the eigenvectors x i , as shown in (8.79). Thus, we haveN∑N∑Ay = A a i x i = a i λ i x i .Alternatively, the action of the RHS of (8.80) on y is given byN∑N∑λ i x i (x i ) † y = a i λ i x i ,i=1i=1since a i =(x i ) † y. We see that the two expressions <strong>for</strong> the action of each side of (8.80) on yare identical, which implies that this relationship is indeed correct. ◭i=1i=18.13.2 Eigenvectors <strong>and</strong> eigenvalues of Hermitian <strong>and</strong> anti-Hermitian matricesFor a normal matrix we showed that if Ax = λx then A † x = λ ∗ x. However, if A isalso Hermitian, A = A † , it follows necessarily that λ = λ ∗ . Thus, the eigenvaluesof an Hermitian matrix are real, a result which may be proved directly.◮Prove that the eigenvalues of an Hermitian matrix are real.For any particular eigenvector x i , we take the Hermitian conjugate of Ax i = λ i x i to give(x i ) † A † = λ ∗ i (x i ) † . (8.81)Using A † = A, sinceA is Hermitian, <strong>and</strong> multiplying on the right by x i ,weobtain(x i ) † Ax i = λ ∗ i (x i ) † x i . (8.82)But multiplying Ax i = λ i x i through on the left by (x i ) † gives(x i ) † Ax i = λ i (x i ) † x i .Subtracting this from (8.82) yields0 =(λ ∗ i − λ i )(x i ) † x i .276


8.13 EIGENVECTORS AND EIGENVALUESBut (x i ) † x i is the modulus squared of the non-zero vector x i <strong>and</strong> is thus non-zero. Henceλ ∗ i must equal λ i <strong>and</strong> thus be real. The same argument can be used to show that theeigenvalues of a real symmetric matrix are themselves real. ◭The importance of the above result will be apparent to any student of quantummechanics. In quantum mechanics the eigenvalues of operators correspond tomeasured values of observable quantities, e.g. energy, angular momentum, parity<strong>and</strong> so on, <strong>and</strong> these clearly must be real. If we use Hermitian operators to<strong>for</strong>mulate the theories of quantum mechanics, the above property guaranteesphysically meaningful results.Since an Hermitian matrix is also a normal matrix, its eigenvectors are orthogonal(or can be made so using the Gram–Schmidt orthogonalisation procedure).Alternatively we can prove the orthogonality of the eigenvectors directly.◮Prove that the eigenvectors corresponding to different eigenvalues of an Hermitian matrixare orthogonal.Consider two unequal eigenvalues λ i <strong>and</strong> λ j <strong>and</strong> their corresponding eigenvectors satisfyingAx i = λ i x i , (8.83)Ax j = λ j x j . (8.84)Taking the Hermitian conjugate of (8.83) we find (x i ) † A † = λ ∗ i (xi ) † . Multiplying this on theright by x j we obtain(x i ) † A † x j = λ ∗ i (x i ) † x j ,<strong>and</strong> similarly multiplying (8.84) through on the left by (x i ) † we find(x i ) † Ax j = λ j (x i ) † x j .Then, since A † = A, the two left-h<strong>and</strong> sides are equal <strong>and</strong>, because the λ i are real, onsubtraction we obtain0 =(λ i − λ j )(x i ) † x j .Finally we note that λ i ≠ λ j <strong>and</strong> so (x i ) † x j = 0, i.e. the eigenvectors x i <strong>and</strong> x j areorthogonal. ◭In the case where some of the eigenvalues are equal, further justification of theorthogonality of the eigenvectors is needed. The Gram–Schmidt orthogonalisationprocedure discussed above provides a proof of, <strong>and</strong> a means of achieving,orthogonality. The general method has already been described <strong>and</strong> we will notrepeat it here.We may also consider the properties of the eigenvalues <strong>and</strong> eigenvectors of ananti-Hermitian matrix, <strong>for</strong> which A † = −A <strong>and</strong> thusAA † = A(−A) =(−A)A = A † A.There<strong>for</strong>e matrices that are anti-Hermitian are also normal <strong>and</strong> so have mutuallyorthogonal eigenvectors. The properties of the eigenvalues are also simplydeduced, since if Ax = λx thenλ ∗ x = A † x = −Ax = −λx.277


MATRICES AND VECTOR SPACESHence λ ∗ = −λ <strong>and</strong> so λ must be pure imaginary (or zero). In a similar mannerto that used <strong>for</strong> Hermitian matrices, these properties may be proved directly.8.13.3 Eigenvectors <strong>and</strong> eigenvalues of a unitary matrixA unitary matrix satisfies A † = A −1 <strong>and</strong> is also a normal matrix, with mutuallyorthogonal eigenvectors. To investigate the eigenvalues of a unitary matrix, wenote that if Ax = λx thenx † x = x † A † Ax = λ ∗ λx † x,<strong>and</strong> we deduce that λλ ∗ = |λ| 2 = 1. Thus, the eigenvalues of a unitary matrixhave unit modulus.8.13.4 Eigenvectors <strong>and</strong> eigenvalues of a general square matrixWhen an N × N matrix is not normal there are no general properties of itseigenvalues <strong>and</strong> eigenvectors; in general it is not possible to find any orthogonalset of N eigenvectors or even to find pairs of orthogonal eigenvectors (exceptby chance in some cases). While the N non-orthogonal eigenvectors are usuallylinearly independent <strong>and</strong> hence <strong>for</strong>m a basis <strong>for</strong> the N-dimensional vector space,this is not necessarily so. It may be shown (although we will not prove it) that anyN × N matrix with distinct eigenvalues has N linearly independent eigenvectors,which there<strong>for</strong>e <strong>for</strong>m a basis <strong>for</strong> the N-dimensional vector space. If a generalsquare matrix has degenerate eigenvalues, however, then it may or may not haveN linearly independent eigenvectors. A matrix whose eigenvectors are not linearlyindependent is said to be defective.8.13.5 Simultaneous eigenvectorsWe may now ask under what conditions two different normal matrices can havea common set of eigenvectors. The result – that they do so if, <strong>and</strong> only if, theycommute – has profound significance <strong>for</strong> the foundations of quantum mechanics.To prove this important result let A <strong>and</strong> B be two N × N normal matrices <strong>and</strong>x i be the ith eigenvector of A corresponding to eigenvalue λ i ,i.e.Ax i = λ i x i <strong>for</strong> i =1, 2,... ,N.For the present we assume that the eigenvalues are all different.(i) First suppose that A <strong>and</strong> B commute. Now considerABx i = BAx i = Bλ i x i = λ i Bx i ,where we have used the commutativity <strong>for</strong> the first equality <strong>and</strong> the eigenvectorproperty <strong>for</strong> the second. It follows that A(Bx i )=λ i (Bx i ) <strong>and</strong> thus that Bx i is an278


8.13 EIGENVECTORS AND EIGENVALUESeigenvector of A corresponding to eigenvalue λ i . But the eigenvector solutions of(A − λ i I)x i = 0 are unique to within a scale factor, <strong>and</strong> we there<strong>for</strong>e conclude thatBx i = µ i x i<strong>for</strong> some scale factor µ i . However, this is just an eigenvector equation <strong>for</strong> B <strong>and</strong>shows that x i is an eigenvector of B, in addition to being an eigenvector of A. Byreversing the roles of A <strong>and</strong> B, it also follows that every eigenvector of B is aneigenvector of A. Thus the two sets of eigenvectors are identical.(ii) Now suppose that A <strong>and</strong> B have all their eigenvectors in common, a typicalone x i satisfying bothAx i = λ i x i <strong>and</strong> Bx i = µ i x i .As the eigenvectors span the N-dimensional vector space, any arbitrary vector xin the space can be written as a linear combination of the eigenvectors,i=1x =N∑c i x i .Now consider bothN∑N∑N∑ABx = AB c i x i = A c i µ i x i = c i λ i µ i x i ,<strong>and</strong>i=1i=1i=1i=1i=1N∑N∑N∑BAx = BA c i x i = B c i λ i x i = c i µ i λ i x i .It follows that ABx <strong>and</strong> BAx are the same <strong>for</strong> any arbitrary x <strong>and</strong> hence that(AB − BA)x = 0<strong>for</strong> all x. Thatis,A <strong>and</strong> B commute.This completes the proof that a necessary <strong>and</strong> sufficient condition <strong>for</strong> twonormal matrices to have a set of eigenvectors in common is that they commute.It should be noted that if an eigenvalue of A, say, is degenerate then not all ofits possible sets of eigenvectors will also constitute a set of eigenvectors of B.However, provided that by taking linear combinations one set of joint eigenvectorscan be found, the proof is still valid <strong>and</strong> the result still holds.When extended to the case of Hermitian operators <strong>and</strong> continuous eigenfunctions(sections 17.2 <strong>and</strong> 17.3) the connection between commuting matrices <strong>and</strong>a set of common eigenvectors plays a fundamental role in the postulatory basisof quantum mechanics. It draws the distinction between commuting <strong>and</strong> noncommutingobservables <strong>and</strong> sets limits on how much in<strong>for</strong>mation about a systemcan be known, even in principle, at any one time.279i=1


MATRICES AND VECTOR SPACES8.14 Determination of eigenvalues <strong>and</strong> eigenvectorsThe next step is to show how the eigenvalues <strong>and</strong> eigenvectors of a given N × Nmatrix A are found. To do this we refer to (8.68) <strong>and</strong> as in (8.69) rewrite it asAx − λIx =(A − λI)x = 0. (8.85)The slight rearrangement used here is to write x as Ix, whereI is the unit matrixof order N. The point of doing this is immediate since (8.85) now has the <strong>for</strong>mof a homogeneous set of simultaneous equations, the theory of which will bedeveloped in section 8.18. What will be proved there is that the equation Bx = 0only has a non-trivial solution x if |B| = 0. Correspondingly, there<strong>for</strong>e, we musthave in the present case that|A − λI| = 0, (8.86)if there are to be non-zero solutions x to (8.85).Equation (8.86) is known as the characteristic equation <strong>for</strong> A <strong>and</strong> its LHS asthe characteristic or secular determinant of A. The equation is a polynomial ofdegree N in the quantity λ. TheN roots of this equation λ i , i =1, 2,...,N, givethe eigenvalues of A. Corresponding to each λ i there will be a column vector x i ,which is the ith eigenvector of A <strong>and</strong> can be found by using (8.68).It will be observed that when (8.86) is written out as a polynomial equation inλ, the coefficient of −λ N−1 in the equation will be simply A 11 + A 22 + ···+ A NNrelative to the coefficient of λ N . As discussed in section 8.8, the quantity ∑ Ni=1 A iiis the trace of A <strong>and</strong>, from the ordinary theory of polynomial equations, will beequal to the sum of the roots of (8.86):N∑λ i =TrA. (8.87)i=1This can be used as one check that a computation of the eigenvalues λ i has beendone correctly. Unless equation (8.87) is satisfied by a computed set of eigenvalues,they have not been calculated correctly. However, that equation (8.87) is satisfied isa necessary, but not sufficient, condition <strong>for</strong> a correct computation. An alternativeproof of (8.87) is given in section 8.16.◮Find the eigenvalues <strong>and</strong> normalised eigenvectors of the real symmetric matrix⎛A = ⎝ 1 1 31 1 −3⎞⎠ .3 −3 −3Using (8.86),∣1 − λ 1 31 1− λ −33 −3 −3 − λ280∣ =0.


8.14 DETERMINATION OF EIGENVALUES AND EIGENVECTORSExp<strong>and</strong>ing out this determinant gives(1 − λ) [(1 − λ)(−3 − λ) − (−3)(−3)] +1[(−3)(3) − 1(−3 − λ)]+3[1(−3) − (1 − λ)(3)] =0,which simplifies to give(1 − λ)(λ 2 +2λ − 12) + (λ − 6) + 3(3λ − 6) = 0,⇒ (λ − 2)(λ − 3)(λ +6)=0.Hence the roots of the characteristic equation, which are the eigenvalues of A, areλ 1 =2,λ 2 =3,λ 3 = −6. We note that, as expected,λ 1 + λ 2 + λ 3 = −1=1+1− 3=A 11 + A 22 + A 33 =TrA.For the first root, λ 1 = 2, a suitable eigenvector x 1 , with elements x 1 , x 2 , x 3 ,mustsatisfyAx 1 =2x 1 or, equivalently,x 1 + x 2 +3x 3 =2x 1 ,x 1 + x 2 − 3x 3 =2x 2 , (8.88)3x 1 − 3x 2 − 3x 3 =2x 3 .These three equations are consistent (to ensure this was the purpose in finding the particularvalues of λ) <strong>and</strong>yieldx 3 =0,x 1 = x 2 = k, wherek is any non-zero number. A suitableeigenvector would thus bex 1 = (k k 0) T .If we apply the normalisation condition, we require k 2 + k 2 +0 2 =1ork =1/ √ 2. Hence( ) T 1x 1 1= √2 √2 0 = √ 1 (1 1 0) T .2Repeating the last paragraph, but with the factor 2 on the RHS of (8.88) replacedsuccessively by λ 2 =3<strong>and</strong>λ 3 = −6, gives two further normalised eigenvectorsx 2 = √ 1 (1 − 1 1) T , x 3 = √ 1 (1 − 1 − 2) T . ◭3 6In the above example, the three values of λ are all different <strong>and</strong> A is areal symmetric matrix. Thus we expect, <strong>and</strong> it is easily checked, that the threeeigenvectors are mutually orthogonal, i.e.(x1 ) Tx 2 = ( x 1) Tx 3 = ( x 2) Tx 3 =0.It will be apparent also that, as expected, the normalisation of the eigenvectorshas no effect on their orthogonality.8.14.1 Degenerate eigenvaluesWe return now to the case of degenerate eigenvalues, i.e. those that have two ormore associated eigenvectors. We have shown already that it is always possibleto construct an orthogonal set of eigenvectors <strong>for</strong> a normal matrix, see subsection8.13.1, <strong>and</strong> the following example illustrates one method <strong>for</strong> constructingsuch a set.281


MATRICES AND VECTOR SPACES◮Construct an orthonormal set of eigenvectors <strong>for</strong> the matrix⎛A = ⎝ 1 0 30 −2 0⎞⎠ .3 0 1We first determine the eigenvalues using |A − λI| =0:1 − λ 0 30=0 −2 − λ 0∣ 3 0 1− λ ∣ = −(1 − λ)2 (2 + λ) + 3(3)(2 + λ)=(4− λ)(λ +2) 2 .Thus λ 1 =4,λ 2 = −2 =λ 3 . The eigenvector x 1 = (x 1 x 2 x 3 ) T is found from⎛⎝ 1 0 3⎞ ⎛0 −2 0 ⎠ ⎝ x ⎞ ⎛1x 2⎠ =4⎝ x ⎞⎛1x 2⎠ ⇒ x 1 = √ 1 ⎝ 1 ⎞0 ⎠ .3 0 1 x 3 x 32 1A general column vector that is orthogonal to x 1 isx = (a b − a) T , (8.89)<strong>and</strong> it is easily shown that⎛⎞ ⎛ ⎞ ⎛ ⎞Ax = ⎝ 1 0 30 −2 0 ⎠ ⎝a b ⎠ = −2 ⎝a b ⎠ = −2x.3 0 1 −a−aThus x is a eigenvector of A with associated eigenvalue −2. It is clear, however, that thereis an infinite set of eigenvectors x all possessing the required property; the geometricalanalogue is that there are an infinite number of corresponding vectors x lying in theplane that has x 1 as its normal. We do require that the two remaining eigenvectors areorthogonal to one another, but this still leaves an infinite number of possibilities. For x 2 ,there<strong>for</strong>e, let us choose a simple <strong>for</strong>m of (8.89), suitably normalised, say,x 2 = (0 1 0) T .The third eigenvector is then specified (to within an arbitrary multiplicative constant)by the requirement that it must be orthogonal to x 1 <strong>and</strong> x 2 ; thus x 3 may be found byevaluating the vector product of x 1 <strong>and</strong> x 2 <strong>and</strong> normalising the result. This givesx 3 = 1 √2(−1 0 1) T ,to complete the construction of an orthonormal set of eigenvectors. ◭8.15 Change of basis <strong>and</strong> similarity trans<strong>for</strong>mationsThroughout this chapter we have considered the vector x as a geometrical quantitythat is independent of any basis (or coordinate system). If we introduce a basise i , i =1, 2,...,N, into our N-dimensional vector space then we may writex = x 1 e 1 + x 2 e 2 + ···+ x N e N ,282


8.15 CHANGE OF BASIS AND SIMILARITY TRANSFORMATIONS<strong>and</strong> represent x in this basis by the column matrixx =(x 1 x 2 ··· x n ) T ,having components x i . We now consider how these components change as a resultof a prescribed change of basis. Let us introduce a new basis e ′ i , i =1, 2,...,N,which is related to the old basis bye ′ j =N∑S ij e i , (8.90)i=1the coefficient S ij being the ith component of e ′ j with respect to the old (unprimed)basis. For an arbitrary vector x it follows thatN∑ N∑N∑x = x i e i = x ′ je ′ j =i=1j=1x ′ jj=1 i=1N∑S ij e i .From this we derive the relationship between the components of x in the twocoordinate systems asN∑x i = S ij x ′ j,which we can write in matrix <strong>for</strong>m asj=1x = Sx ′ (8.91)where S is the trans<strong>for</strong>mation matrix associated with the change of basis.Furthermore, since the vectors e ′ j are linearly independent, the matrix S isnon-singular <strong>and</strong> so possesses an inverse S −1 . Multiplying (8.91) on the left byS −1 we findx ′ = S −1 x, (8.92)which relates the components of x in the new basis to those in the old basis.Comparing (8.92) <strong>and</strong> (8.90) we note that the components of x trans<strong>for</strong>m inverselyto the way in which the basis vectors e i themselves trans<strong>for</strong>m. This has to be so,as the vector x itself must remain unchanged.We may also find the trans<strong>for</strong>mation law <strong>for</strong> the components of a linearoperator under the same change of basis. Now, the operator equation y = A x(which is basis independent) can be written as a matrix equation in each of thetwo bases asBut, using (8.91), we may rewrite the first equation asy = Ax, y ′ = A ′ x ′ . (8.93)Sy ′ = ASx ′ ⇒ y ′ = S −1 ASx ′ .283


MATRICES AND VECTOR SPACESComparing this with the second equation in (8.93) we find that the componentsof the linear operator A trans<strong>for</strong>m asA ′ = S −1 AS. (8.94)Equation (8.94) is an example of a similarity trans<strong>for</strong>mation – a trans<strong>for</strong>mationthat can be particularly useful in converting matrices into convenient <strong>for</strong>ms <strong>for</strong>computation.Given a square matrix A, we may interpret it as representing a linear operatorA in a given basis e i . From (8.94), however, we may also consider the matrixA ′ = S −1 AS, <strong>for</strong> any non-singular matrix S, as representing the same linearoperator A but in a new basis e ′ j , related to the old basis bye ′ j = ∑ iS ij e i .There<strong>for</strong>e we would expect that any property of the matrix A that representssome (basis-independent) property of the linear operator A will also be sharedby the matrix A ′ . We list these properties below.(i) If A = I then A ′ = I, since, from (8.94),(ii) The value of the determinant is unchanged:A ′ = S −1 IS = S −1 S = I. (8.95)|A ′ | = |S −1 AS| = |S −1 ||A||S| = |A||S −1 ||S| = |A||S −1 S| = |A|. (8.96)(iii) The characteristic determinant <strong>and</strong> hence the eigenvalues of A ′ are thesame as those of A: from (8.86),|A ′ − λI| = |S −1 AS − λI| = |S −1 (A − λI)S|= |S −1 ||S||A − λI| = |A − λI|. (8.97)(iv) The value of the trace is unchanged: from (8.87),Tr A ′ = ∑ A ′ ii = ∑ ∑ ∑(S −1 ) ij A jk S kiii j k= ∑ ∑ ∑S ki (S −1 ) ij A jk = ∑ ∑δ kj A jk = ∑ A jji j kj kj=TrA. (8.98)An important class of similarity trans<strong>for</strong>mations is that <strong>for</strong> which S is a unitarymatrix; in this case A ′ = S −1 AS = S † AS. Unitary trans<strong>for</strong>mation matricesare particularly important, <strong>for</strong> the following reason. If the original basis e i is284


8.16 DIAGONALISATION OF MATRICESorthonormal <strong>and</strong> the trans<strong>for</strong>mation matrix S is unitary then〈∑ ∣〈e ′ i|e ′ ∣∣∑ 〉j〉 = S ki e k S rj e r= ∑ k= ∑ kkS ∗ kiS ∗ kir∑S rj 〈e k |e r 〉r∑S rj δ kr = ∑ SkiS ∗ kj =(S † S) ij = δ ij ,rkshowing that the new basis is also orthonormal.Furthermore, in addition to the properties of general similarity trans<strong>for</strong>mations,<strong>for</strong> unitary trans<strong>for</strong>mations the following hold.(i) If A is Hermitian (anti-Hermitian) then A ′ is Hermitian (anti-Hermitian),i.e. if A † = ±A then(A ′ ) † =(S † AS) † = S † A † S = ±S † AS = ±A ′ . (8.99)(ii) If A is unitary (so that A † = A −1 )thenA ′ is unitary, since(A ′ ) † A ′ =(S † AS) † (S † AS) =S † A † SS † AS = S † A † AS= S † IS = I. (8.100)8.16 Diagonalisation of matricesSuppose that a linear operator A is represented in some basis e i , i =1, 2,...,N,by the matrix A. Consider a new basis x j given byx j =N∑S ij e i ,where the x j are chosen to be the eigenvectors of the linear operator A ,i.e.i=1A x j = λ j x j . (8.101)In the new basis, A is represented by the matrix A ′ = S −1 AS, which has aparticularly simple <strong>for</strong>m, as we shall see shortly. The element S ij of S is the ithcomponent, in the old (unprimed) basis, of the jth eigenvector x j of A, i.e.thecolumns of S are the eigenvectors of the matrix A:⎛S = ⎝↑ ↑ ↑x 1 x 2 ··· x N↓ ↓ ↓285⎞⎠ ,


MATRICES AND VECTOR SPACESthat is, S ij =(x j ) i . There<strong>for</strong>e A ′ is given by(S −1 AS) ij = ∑ ∑(S −1 ) ik A kl S ljk l= ∑ ∑(S −1 ) ik A kl (x j ) lk l= ∑ k= ∑ k(S −1 ) ik λ j (x j ) kλ j (S −1 ) ik S kj = λ j δ ij .So the matrix A ′ is diagonal with the eigenvalues of A as the diagonal elements,i.e.⎛λ 1 0 ··· 0⎞0 ··· 0 λ N A ′ =0 λ 2 .⎜⎝. ⎟. .. 0 ⎠ .There<strong>for</strong>e, given a matrix A, if we construct the matrix S that has the eigenvectorsof A as its columns then the matrix A ′ = S −1 AS is diagonal <strong>and</strong> has theeigenvalues of A as its diagonal elements. Since we require S to be non-singular(|S| ≠ 0), the N eigenvectors of A must be linearly independent <strong>and</strong> <strong>for</strong>m a basis<strong>for</strong> the N-dimensional vector space. It may be shown that any matrix with distincteigenvalues can be diagonalised by this procedure. If, however, a general squarematrix has degenerate eigenvalues then it may, or may not, have N linearlyindependent eigenvectors. If it does not then it cannot be diagonalised.For normal matrices (which include Hermitian, anti-Hermitian <strong>and</strong> unitarymatrices) the N eigenvectors are indeed linearly independent. Moreover, whennormalised, these eigenvectors <strong>for</strong>m an orthonormal set (or can be made to doso). There<strong>for</strong>e the matrix S with these normalised eigenvectors as columns, i.e.whose elements are S ij =(x j ) i ,hastheproperty(S † S) ij = ∑ k(S † ) ik (S) kj = ∑ kS ∗ kiS kj = ∑ k(x i ) ∗ k(x j ) k = (x i ) † x j = δ ij .Hence S is unitary (S −1 = S † ) <strong>and</strong> the original matrix A can be diagonalised byA ′ = S −1 AS = S † AS.There<strong>for</strong>e, any normal matrix A can be diagonalised by a similarity trans<strong>for</strong>mationusing a unitary trans<strong>for</strong>mation matrix S.286


8.16 DIAGONALISATION OF MATRICES◮Diagonalise the matrixA =⎛⎝ 1 0 0 −2 30⎞⎠ .3 0 1The matrix A is symmetric <strong>and</strong> so may be diagonalised by a trans<strong>for</strong>mation of the <strong>for</strong>mA ′ = S † AS, whereS has the normalised eigenvectors of A as its columns. We have alreadyfound these eigenvectors in subsection 8.14.1, <strong>and</strong> so we can write straightaway⎛⎞S = √ 1 1√0 −1⎝ 0 2 0 ⎠ .2 1 0 1We note that although the eigenvalues of A are degenerate, its three eigenvectors arelinearly independent <strong>and</strong> so A can still be diagonalised. Thus, calculating S † AS we obtain⎛⎞ ⎛S † AS = 1 1√0 1⎝ 0 2 0 ⎠ ⎝ 1 0 3⎞ ⎛⎞1√0 −10 −2 0 ⎠ ⎝ 0 2 0 ⎠2−1 0 1 3 0 1 1 0 1⎛= ⎝ 4 0 0⎞0 −2 0 ⎠ ,0 0 −2which is diagonal, as required, <strong>and</strong> has as its diagonal elements the eigenvalues of A. ◭If a matrix A is diagonalised by the similarity trans<strong>for</strong>mation A ′ = S −1 AS, sothat A ′ = diag(λ 1 ,λ 2 ,...,λ N ), then we have immediatelyTr A ′ =TrA =N∑λ i , (8.102)i=1|A ′ | = |A| =N∏λ i , (8.103)since the eigenvalues of the matrix are unchanged by the trans<strong>for</strong>mation. Moreover,these results may be used to prove the rather useful trace <strong>for</strong>mulai=1| exp A| = exp(Tr A), (8.104)where the exponential of a matrix is as defined in (8.38).◮Prove the trace <strong>for</strong>mula (8.104).At the outset, we note that <strong>for</strong> the similarity trans<strong>for</strong>mation A ′ = S −1 AS, we have(A ′ ) n =(S −1 AS)(S −1 AS) ···(S −1 AS) =S −1 A n S.Thus, from (8.38), we obtain exp A ′ = S −1 (exp A)S, from which it follows that | exp A ′ | =287


MATRICES AND VECTOR SPACES| exp A|. Moreover, by choosing the similarity trans<strong>for</strong>mation so that it diagonalises A, wehave A ′ =diag(λ 1 ,λ 2 ,...,λ N ), <strong>and</strong> soN∏| exp A| = | exp A ′ | = | exp[diag(λ 1 ,λ 2 ,...,λ N )]| = |diag(exp λ 1 , exp λ 2 ,...,exp λ N )| = exp λ i .Rewriting the final product of exponentials of the eigenvalues as the exponential of thesum of the eigenvalues, we find(N∏N)∑| exp A| = exp λ i =exp λ i =exp(TrA),i=1which gives the trace <strong>for</strong>mula (8.104). ◭i=1i=18.17 Quadratic <strong>and</strong> Hermitian <strong>for</strong>msLet us now introduce the concept of quadratic <strong>for</strong>ms (<strong>and</strong> their complex analogues,Hermitian <strong>for</strong>ms). A quadratic <strong>for</strong>m Q is a scalar function of a real vectorx given byQ(x) =〈x|A x〉, (8.105)<strong>for</strong> some real linear operator A . In any given basis (coordinate system) we canwrite (8.105) in matrix <strong>for</strong>m asQ(x) =x T Ax, (8.106)where A is a real matrix. In fact, as will be explained below, we need only considerthe case where A is symmetric, i.e. A = A T . As an example in a three-dimensionalspace,( ) ⎛ 1 1 3Q = x T Ax = x 1 x 2 x 3⎝ 1 1 −33 −3 −3⎞ ⎛⎠ ⎝x 1x 2x 3⎞⎠= x 2 1 + x 2 2 − 3x 2 3 +2x 1 x 2 +6x 1 x 3 − 6x 2 x 3 . (8.107)It is reasonable to ask whether a quadratic <strong>for</strong>m Q = x T Mx, whereM is any(possibly non-symmetric) real square matrix, is a more general definition. Thatthis is not the case may be seen by expressing M in terms of a symmetric matrixA = 1 2 (M+MT ) <strong>and</strong> an antisymmetric matrix B = 1 2 (M−MT ) such that M = A+B.We then haveHowever, Q is a scalar quantity <strong>and</strong> soQ = x T Mx = x T Ax + x T Bx. (8.108)Q = Q T =(x T Ax) T +(x T Bx) T = x T A T x + x T B T x = x T Ax − x T Bx.(8.109)Comparing (8.108) <strong>and</strong> (8.109) shows that x T Bx = 0, <strong>and</strong> hence x T Mx = x T Ax,288


8.17 QUADRATIC AND HERMITIAN FORMSi.e. Q is unchanged by considering only the symmetric part of M. Hence, with noloss of generality, we may assume A = A T in (8.106).From its definition (8.105), Q is clearly a basis- (i.e. coordinate-) independentquantity. Let us there<strong>for</strong>e consider a new basis related to the old one by anorthogonal trans<strong>for</strong>mation matrix S, the components in the two bases of anyvector x being related (as in (8.91)) by x = Sx ′ or, equivalently, by x ′ = S −1 x =S T x. We then haveQ = x T Ax =(x ′ ) T S T ASx ′ =(x ′ ) T A ′ x ′ ,where (as expected) the matrix describing the linear operator A in the newbasis is given by A ′ = S T AS (since S T = S −1 ). But, from the last section, if wechoose as S the matrix whose columns are the normalised eigenvectors of A thenA ′ = S T AS is diagonal with the eigenvalues of A as the diagonal elements. (SinceA is symmetric, its normalised eigenvectors are orthogonal, or can be made so,<strong>and</strong> hence S is orthogonal with S −1 = S T .)In the new basisQ = x T Ax =(x ′ ) T Λx ′ = λ 1 x ′ 12 + λ 2 x ′ 22 + ···+ λ N x ′ N2 , (8.110)where Λ = diag(λ 1 ,λ 2 ,...,λ N )<strong>and</strong>theλ i are the eigenvalues of A. It should benoted that Q contains no cross-terms of the <strong>for</strong>m x ′ 1 x′ 2 .◮Find an orthogonal trans<strong>for</strong>mation that takes the quadratic <strong>for</strong>m (8.107) into the <strong>for</strong>mλ 1 x ′ 12 + λ 2 x ′ 22 + λ 3 x ′ 32 .The required trans<strong>for</strong>mation matrix S has the normalised eigenvectors of A as its columns.We have already found these in section 8.14, <strong>and</strong> so we can write immediately⎛ √ √ ⎞3S = √ 1 √ √ 2 1⎝ 3 − 2 −1 ⎠6 √,0 2 −2which is easily verified as being orthogonal. Since the eigenvalues of A are λ =2,3,<strong>and</strong>−6, the general result already proved shows that the trans<strong>for</strong>mation x = Sx ′ will carry(8.107) into the <strong>for</strong>m 2x ′ 12 +3x ′ 22 − 6x ′ 32 . This may be verified most easily by writing outthe inverse trans<strong>for</strong>mation x ′ = S −1 x = S T x <strong>and</strong> substituting. The inverse equations arex ′ 1 =(x 1 + x 2 )/ √ 2,x ′ 2 =(x 1 − x 2 + x 3 )/ √ 3, (8.111)x ′ 3 =(x 1 − x 2 − 2x 3 )/ √ 6.If these are substituted into the <strong>for</strong>m Q =2x ′ 12 +3x ′ 22 − 6x ′ 32 then the original expression(8.107) is recovered. ◭In the definition of Q it was assumed that the components x 1 , x 2 , x 3 <strong>and</strong> thematrix A were real. It is clear that in this case the quadratic <strong>for</strong>m Q ≡ x T Ax is real289


MATRICES AND VECTOR SPACESalso. Another, rather more general, expression that is also real is the Hermitian<strong>for</strong>mH(x) ≡ x † Ax, (8.112)where A is Hermitian (i.e. A † = A) <strong>and</strong> the components of x may now be complex.It is straight<strong>for</strong>ward to show that H is real, sinceH ∗ =(H T ) ∗ = x † A † x = x † Ax = H.With suitable generalisation, the properties of quadratic <strong>for</strong>ms apply also to Hermitian<strong>for</strong>ms, but to keep the presentation simple we will restrict our discussionto quadratic <strong>for</strong>ms.A special case of a quadratic (Hermitian) <strong>for</strong>m is one <strong>for</strong> which Q = x T Axis greater than zero <strong>for</strong> all column matrices x. By choosing as the basis theeigenvectors of A we have Q in the <strong>for</strong>mQ = λ 1 x 2 1 + λ 2 x 2 2 + λ 3 x 2 3.The requirement that Q>0 <strong>for</strong> all x means that all the eigenvalues λ i of A mustbe positive. A symmetric (Hermitian) matrix A with this property is called positivedefinite. If,instead,Q ≥ 0 <strong>for</strong> all x then it is possible that some of the eigenvaluesare zero, <strong>and</strong> A is called positive semi-definite.8.17.1 The stationary properties of the eigenvectorsConsider a quadratic <strong>for</strong>m, such as Q(x) =〈x|A x〉, equation (8.105), in a fixedbasis. As the vector x is varied, through changes in its three components x 1 , x 2<strong>and</strong> x 3 , the value of the quantity Q also varies. Because of the homogeneous<strong>for</strong>m of Q we may restrict any investigation of these variations to vectors of unitlength (since multiplying any vector x by any scalar k simply multiplies the valueof Q by a factor k 2 ).Of particular interest are any vectors x that make the value of the quadratic<strong>for</strong>m a maximum or minimum. A necessary, but not sufficient, condition <strong>for</strong> thisis that Q is stationary with respect to small variations ∆x in x, whilst 〈x|x〉 ismaintained at a constant value (unity).In the chosen basis the quadratic <strong>for</strong>m is given by Q = x T Ax <strong>and</strong>, usingLagrange undetermined multipliers to incorporate the variational constraints, weare led to seek solutions of∆[x T Ax − λ(x T x − 1)] = 0. (8.113)This may be used directly, together with the fact that (∆x T )Ax = x T A ∆x, sinceAis symmetric, to obtainAx = λx (8.114)290


8.17 QUADRATIC AND HERMITIAN FORMSas the necessary condition that x must satisfy. If (8.114) is satisfied <strong>for</strong> someeigenvector x then the value of Q(x) is given byQ = x T Ax = x T λx = λ. (8.115)However, if x <strong>and</strong> y are eigenvectors corresponding to different eigenvalues thenthey are (or can be chosen to be) orthogonal. Consequently the expression y T Axis necessarily zero, sincey T Ax = y T λx = λy T x =0. (8.116)Summarising, those column matrices x of unit magnitude that make thequadratic <strong>for</strong>m Q stationary are eigenvectors of the matrix A, <strong>and</strong> the stationaryvalue of Q is then equal to the corresponding eigenvalue. It is straight<strong>for</strong>wardto see from the proof of (8.114) that, conversely, any eigenvector of A makes Qstationary.Instead of maximising or minimising Q = x T Ax subject to the constraintx T x = 1, an equivalent procedure is to extremise the functionλ(x) = xT Axx T x .◮Show that if λ(x) is stationary then x is an eigenvector of A <strong>and</strong> λ(x) is equal to thecorresponding eigenvalue.We require ∆λ(x) = 0 with respect to small variations in x. Now∆λ = 1 [(x T x) ( ∆x T Ax + x T A ∆x ) − x T Ax ( ∆x T x + x T ∆x )](x T x) 2 ( )= 2∆xT Ax x T Ax ∆x T x− 2x T x x T x x T x ,since x T A ∆x =(∆x T )Ax <strong>and</strong> x T ∆x =(∆x T )x. Thus∆λ = 2x T x ∆xT [Ax − λ(x)x].Hence, if ∆λ =0thenAx = λ(x)x, i.e.x is an eigenvector of A with eigenvalue λ(x). ◭Thus the eigenvalues of a symmetric matrix A are the values of the functionλ(x) = xT Axx T xat its stationary points. The eigenvectors of A lie along those directions in space<strong>for</strong> which the quadratic <strong>for</strong>m Q = x T Ax has stationary values, given a fixedmagnitude <strong>for</strong> the vector x. Similar results hold <strong>for</strong> Hermitian matrices.291


MATRICES AND VECTOR SPACES8.17.2 Quadratic surfacesThe results of the previous subsection may be turned round to state that thesurface given byx T Ax = constant = 1 (say) (8.117)<strong>and</strong> called a quadratic surface, has stationary values of its radius (i.e. origin–surface distance) in those directions that are along the eigenvectors of A. Morespecifically, in three dimensions the quadratic surface x T Ax = 1 has its principalaxes along the three mutually perpendicular eigenvectors of A, <strong>and</strong>thesquaresof the corresponding principal radii are given by λ −1i , i =1, 2, 3. As well ashaving this stationary property of the radius, a principal axis is characterised bythe fact that any section of the surface perpendicular to it has some degree ofsymmetry about it. If the eigenvalues corresponding to any two principal axes aredegenerate then the quadratic surface has rotational symmetry about the thirdprincipal axis <strong>and</strong> the choice of a pair of axes perpendicular to that axis is notuniquely defined.◮Find the shape of the quadratic surfacex 2 1 + x 2 2 − 3x 2 3 +2x 1 x 2 +6x 1 x 3 − 6x 2 x 3 =1.If, instead of expressing the quadratic surface in terms of x 1 , x 2 , x 3 , as in (8.107), wewere to use the new variables x ′ 1 , x′ 2 , x′ 3 defined in (8.111), <strong>for</strong> which the coordinate axesare along the three mutually perpendicular eigenvector directions (1, 1, 0), (1, −1, 1) <strong>and</strong>(1, −1, −2), then the equation of the surface would take the <strong>for</strong>m (see (8.110))x ′ 212(1/ √ 2) + x′ 22 (1/ √ 3) − x′ 32 (1/ √ 6) =1.2Thus, <strong>for</strong> example, a section of the quadratic surface in the plane x ′ 3 =0,i.e.x 1 − x 2 −2x 3 = 0, is an ellipse, with semi-axes 1/ √ 2<strong>and</strong>1/ √ 3. Similarly a section in the planex ′ 1 = x 1 + x 2 = 0 is a hyperbola. ◭Clearly the simplest three-dimensional situation to visualise is that in which allthe eigenvalues are positive, since then the quadratic surface is an ellipsoid.28.18 Simultaneous linear equationsIn physical applications we often encounter sets of simultaneous linear equations.In general we may have M equations in N unknowns x 1 ,x 2 ,...,x N of the <strong>for</strong>mA 11 x 1 + A 12 x 2 + ···+ A 1N x N = b 1 ,A 21 x 1 + A 22 x 2 + ···+ A 2N x N = b 2 ,..A M1 x 1 + A M2 x 2 + ···+ A MN x N = b M ,(8.118)292


8.18 SIMULTANEOUS LINEAR EQUATIONSwhere the A ij <strong>and</strong> b i have known values. If all the b i are zero then the system ofequations is called homogeneous, otherwise it is inhomogeneous. Depending on thegiven values, this set of equations <strong>for</strong> the N unknowns x 1 , x 2 , ..., x N may haveeither a unique solution, no solution or infinitely many solutions. Matrix analysismay be used to distinguish between the possibilities. The set of equations may beexpressed as a single matrix equation Ax = b, or, written out in full, as⎛ ⎞⎛⎞ ⎛ ⎞ bA 11 A 12 ... A 1N x 11A 21 A 22 ... A 2Nx 2b ⎜⎝. ⎟ ⎜ ⎟. . .. . ⎠ ⎝ . ⎠ = 2.⎜⎝ . ⎟⎠A M1 A M2 ... A MNx Nb M8.18.1 The range <strong>and</strong> null space of a matrixAs we discussed in section 8.2, we may interpret the matrix equation Ax = b asrepresenting, in some basis, the linear trans<strong>for</strong>mation A x = b of a vector x in anN-dimensional vector space V into a vector b in some other (in general different)M-dimensional vector space W .In general the operator A will map any vector in V into some particularsubspace of W , which may be the entire space. This subspace is called the rangeof A (or A) <strong>and</strong> its dimension is equal to the rank of A. Moreover, if A (<strong>and</strong>hence A) issingular then there exists some subspace of V that is mapped ontothe zero vector 0 in W ; that is, any vector y that lies in the subspace satisfiesA y = 0. This subspace is called the null space of A <strong>and</strong> the dimension of thisnull space is called the nullity of A. We note that the matrix A must be singularif M ≠ N <strong>and</strong> may be singular even if M = N.The dimensions of the range <strong>and</strong> the null space of a matrix are related throughthe fundamental relationshiprank A + nullity A = N, (8.119)where N is the number of original unknowns x 1 ,x 2 ,...,x N .◮Prove the relationship (8.119).As discussed in section 8.11, if the columns of an M × N matrix A are interpreted as thecomponents, in a given basis, of N (M-component) vectors v 1 , v 2 ,...,v N then rank A isequal to the number of linearly independent vectors in this set (this number is also equalto the dimension of the vector space spanned by these vectors). Writing (8.118) in termsof the vectors v 1 , v 2 ,...,v N , we havex 1 v 1 + x 2 v 2 + ···+ x N v N = b. (8.120)From this expression, we immediately deduce that the range of A is merely the span ofthe vectors v 1 , v 2 ,...,v N <strong>and</strong> hence has dimension r =rankA.293


MATRICES AND VECTOR SPACESIf a vector y lies in the null space of A then A y = 0, whichwemaywriteasy 1 v 1 + y 2 v 2 + ···+ y N v N = 0. (8.121)As just shown above, however, only r (≤ N) of these vectors are linearly independent. Byrenumbering, if necessary, we may assume that v 1 , v 2 ,...,v r <strong>for</strong>m a linearly independentset; the remaining vectors, v r+1 , v r+2 ,...,v N , can then be written as a linear superpositionof v 1 , v 2 ,...,v r . We are there<strong>for</strong>e free to choose the N − r coefficients y r+1 ,y r+2 ,...,y Narbitrarily <strong>and</strong> (8.121) will still be satisfied <strong>for</strong> some set of r coefficients y 1 ,y 2 ,...,y r (whichare not all zero). The dimension of the null space is there<strong>for</strong>e N − r, <strong>and</strong> this completesthe proof of (8.119). ◭Equation (8.119) has far-reaching consequences <strong>for</strong> the existence of solutionsto sets of simultaneous linear equations such as (8.118). As mentioned previously,these equations may have no solution, aunique solution or infinitely many solutions.We now discuss these three cases in turn.No solutionThe system of equations possesses no solution unless b lies in the range of A ;inthis case (8.120) will be satisfied <strong>for</strong> some x 1 ,x 2 ,...,x N . This in turn requires thesetofvectorsb, v 1 , v 2 ,...,v N to have the same span (see (8.8)) as v 1 , v 2 ,...,v N .Interms of matrices, this is equivalent to the requirement that the matrix A <strong>and</strong> theaugmented matrix⎛⎞A 11 A 12 ... A 1N b 1A 21 A 22 ... A 2N b 1M = ⎜⎝. ⎟... . ⎠A M1 A M2 ... A MN b Mhave the same rank r. If this condition is satisfied then b does lie in the range ofA , <strong>and</strong> the set of equations (8.118) will have either a unique solution or infinitelymany solutions. If, however, A <strong>and</strong> M have different ranks then there will be nosolution.A unique solutionIf b lies in the range of A <strong>and</strong> if r = N then all the vectors v 1 , v 2 ,...,v N in (8.120)are linearly independent <strong>and</strong> the equation has a unique solution x 1 ,x 2 ,...,x N .Infinitely many solutionsIf b lies in the range of A <strong>and</strong> if r


8.18 SIMULTANEOUS LINEAR EQUATIONSany vector in the null space of A (i.e. A y = 0) thenA (x + y) =A x + A y = A x + 0 = b,<strong>and</strong> so x + y is also a solution. Since the null space is (n − r)-dimensional, so toois the space of solutions.We may use the above results to investigate the special case of the solution ofa homogeneous set of linear equations, <strong>for</strong> which b = 0. Clearly the set always hasthe trivial solution x 1 = x 2 = ··· = x n =0,<strong>and</strong>ifr = N this will be the onlysolution. If r


MATRICES AND VECTOR SPACES◮Show that the set of simultaneous equations2x 1 +4x 2 +3x 3 =4,x 1 − 2x 2 − 2x 3 =0, (8.123)−3x 1 +3x 2 +2x 3 = −7,has a unique solution, <strong>and</strong> find that solution.The simultaneous equations can be represented by the matrix equation Ax = b, i.e.⎛⎝ 2 4 3⎞ ⎛1 −2 −2 ⎠ ⎝ x ⎞ ⎛1x 2⎠ = ⎝ 4 ⎞0 ⎠ .−3 3 2 x 3 −7As we have already shown that A −1 exists <strong>and</strong> have calculated it, see (8.59), it follows thatx = A −1 b or, more explicitly, that⎛ ⎞ ⎛⎞ ⎛ ⎞ ⎛ ⎞⎝ x 1x 2⎠ = 1 ⎝ 2 1 −24 13 7 ⎠ ⎝ 4 0 ⎠ = ⎝ 2−3x113−3 −18 −8 −7 4⎠ . (8.124)Thus the unique solution is x 1 =2,x 2 = −3, x 3 =4.◭LU decompositionAlthough conceptually simple, finding the solution by calculating A −1 can becomputationally dem<strong>and</strong>ing, especially when N is large. In fact, as we shall nowshow, it is not necessary to per<strong>for</strong>m the full inversion of A in order to solve thesimultaneous equations Ax = b. Rather, we can per<strong>for</strong>m a decomposition of thematrix into the product of a square lower triangular matrix L <strong>and</strong> a square uppertriangular matrix U, which are such thatA = LU, (8.125)<strong>and</strong> then use the fact that triangular systems of equations can be solved verysimply.We must begin, there<strong>for</strong>e, by finding the matrices L <strong>and</strong> U such that (8.125)is satisfied. This may be achieved straight<strong>for</strong>wardly by writing out (8.125) incomponent <strong>for</strong>m. For illustration, let us consider the 3 × 3 case. It is, in fact,always possible, <strong>and</strong> convenient, to take the diagonal elements of L as unity, sowe have⎛A = ⎝⎛= ⎝1 0 0L 21 1 0L 31 L 32 1⎞ ⎛⎠ ⎝U 11 U 12 U 13⎞0 U 22 U 23⎠U 11 U 12 U 13⎞L 21 U 11 L 21 U 12 + U 22 L 21 U 13 + U 23⎠ (8.126)The nine unknown elements of L <strong>and</strong> U can now be determined by equating296


8.18 SIMULTANEOUS LINEAR EQUATIONSthe nine elements of (8.126) to those of the 3 × 3 matrix A. This is done in theparticular order illustrated in the example below.Once the matrices L <strong>and</strong> U have been determined, one can use the decompositionto solve the set of equations Ax = b in the following way. From (8.125), we haveLUx = b, but this can be written as two triangular sets of equationsLy = b <strong>and</strong> Ux = y,where y is another column matrix to be determined. One may easily solve the firsttriangular set of equations <strong>for</strong> y, which is then substituted into the second set.The required solution x is then obtained readily from the second triangular setof equations. We note that, as with direct inversion, once the LU decompositionhas been determined, one can solve <strong>for</strong> various RHS column matrices b 1 , b 2 , ... ,with little extra work.◮Use LU decomposition to solve the set of simultaneous equations (8.123).We begin the determination of the matrices L <strong>and</strong> U by equating the elements of thematrix in (8.126) with those of the matrix⎛A = 1 −2 −2⎝ 2 4 3⎞⎠ .−3 3 2This is per<strong>for</strong>med in the following order:1st row: U 11 =2, U 12 =4, U 13 =31st column: L 21 U 11 =1, L 31 U 11 = −3 ⇒ L 21 = 1 , L 2 31 = − 3 22nd row: L 21 U 12 + U 22 = −2 L 21 U 13 + U 23 = −2 ⇒ U 22 = −4, U 23 = − 7 22nd column: L 31 U 12 + L 32 U 22 =3 ⇒ L 32 = − 9 43rd row: L 31 U 13 + L 32 U 23 + U 33 =2 ⇒ U 33 = − 118Thus we may write the matrix A as⎛1 0 0⎞ ⎛2 4 3⎞⎜ 1A = LU = ⎝ 21 0⎟ ⎜⎠ ⎝ 0 −4 − 7 2⎟⎠ .− 3 2− 9 41 0 0 − 118We must now solve the set of equations Ly = b, whichread⎛⎜⎝⎞ ⎛1 0 01⎟ ⎜1 02 ⎠ ⎝− 3 − 9 12 4y 1y 2y 3⎞ ⎛⎟ ⎜⎠ = ⎝Since this set of equations is triangular, we quickly findy 1 =4, y 2 =0− ( 1 )(4) = −2, y 2 3 = −7 − (− 3 )(4) − (− 9 11)(−2) = − .2 4 2These values must then be substituted into the equations Ux = y, whichread⎛⎞ ⎛ ⎞ ⎛ ⎞2 4 3 x 1 4⎜⎝ 0 −4 − 7 ⎟ ⎜ ⎟ ⎜2 ⎠ ⎝ x 2 ⎠ = ⎝ −2⎟⎠ .0 0 − 11 x83 − 11229740−7⎞⎟⎠ .


MATRICES AND VECTOR SPACESThis set of equations is also triangular, <strong>and</strong> we easily find the solutionx 1 =2, x 2 = −3, x 3 =4,which agrees with the result found above by direct inversion. ◭We note, in passing, that one can calculate both the inverse <strong>and</strong> the determinantof A from its LU decomposition. To find the inverse A −1 , one solves the systemof equations Ax = b repeatedly <strong>for</strong> the N different RHS column matrices b = e i ,i =1, 2,...,N,wheree i is the column matrix with its ith element equal to unity<strong>and</strong> the others equal to zero. The solution x in each case gives the correspondingcolumn of A −1 . Evaluation of the determinant |A| is much simpler. From (8.125),we have|A| = |LU| = |L||U|. (8.127)Since L <strong>and</strong> U are triangular, however, we see from (8.64) that their determinantsare equal to the products of their diagonal elements. Since L ii = 1 <strong>for</strong> all i, wethus findN∏|A| = U 11 U 22 ···U NN = U ii .As an illustration, in the above example we find |A| = (2)(−4)(−11/8) = 11,which, as it must, agrees with our earlier calculation (8.58).Finally, we note that if the matrix A is symmetric <strong>and</strong> positive semi-definitethen we can decompose it asi=1A = LL † , (8.128)where L is a lower triangular matrix whose diagonal elements are not, in general,equal to unity. This is known as a Cholesky decomposition (in the special casewhere A is real, the decomposition becomes A = LL T ). The reason that we cannotset the diagonal elements of L equal to unity in this case is that we require thesame number of independent elements in L as in A. The requirement that thematrix be positive semi-definite is easily derived by considering the Hermitian<strong>for</strong>m (or quadratic <strong>for</strong>m in the real case)x † Ax = x † LL † x =(L † x) † (L † x).Denoting the column matrix L † x by y, we see that the last term on the RHSis y † y, which must be greater than or equal to zero. Thus, we require x † Ax ≥ 0<strong>for</strong> any arbitrary column matrix x, <strong>and</strong>soA must be positive semi-definite (seesection 8.17).We recall that the requirement that a matrix be positive semi-definite is equivalentto dem<strong>and</strong>ing that all the eigenvalues of A are positive or zero. If one ofthe eigenvalues of A is zero, however, then from (8.103) we have |A| = 0 <strong>and</strong> so Ais singular. Thus, if A is a non-singular matrix, it must be positive definite (rather298


8.18 SIMULTANEOUS LINEAR EQUATIONSthan just positive semi-definite) in order to per<strong>for</strong>m the Cholesky decomposition(8.128). In fact, in this case, the inability to find a matrix L that satisfies (8.128)implies that A cannot be positive definite.The Cholesky decomposition can be applied in an analogous way to the LUdecomposition discussed above, but we shall not explore it further.Cramer’s ruleAn alternative method of solution is to use Cramer’s rule, which also providessome insight into the nature of the solutions in the various cases. To illustratethis method let us consider a set of three equations in three unknowns,A 11 x 1 + A 12 x 2 + A 13 x 3 = b 1 ,A 21 x 1 + A 22 x 2 + A 23 x 3 = b 2 , (8.129)A 31 x 1 + A 32 x 2 + A 33 x 3 = b 3 ,which may be represented by the matrix equation Ax = b. We wish either to findthe solution(s) x to these equations or to establish that there are no solutions.From result (vi) of subsection 8.9.1, the determinant |A| is unchanged by addingto its first column the combinationx 2× (second column of |A|)+ x 3× (third column of |A|).x 1 x 1We thus obtain∣ A 11 A 12 A 13 ∣∣∣∣∣ |A| =A 21 A 22 A 23 =∣ A 31 A 32 A ∣ 33∣A 11 +(x 2 /x 1 )A 12 +(x 3 /x 1 )A 13 A 12 A 13 ∣∣∣∣∣A 21 +(x 2 /x 1 )A 22 +(x 3 /x 1 )A 23 A 22 A 23 ,A 31 +(x 2 /x 1 )A 32 +(x 3 /x 1 )A 33 A 32 A 33which, on substituting b i /x 1 <strong>for</strong> the ith entry in the first column, yields|A| = 1 x 1∣ ∣∣∣∣∣ b 1 A 12 A 13b 2 A 22 A 23b 3 A 32 A 33∣ ∣∣∣∣∣= 1 x 1∆ 1 .The determinant ∆ 1 is known as a Cramer determinant. Similar manipulations ofthe second <strong>and</strong> third columns of |A| yield x 2 <strong>and</strong> x 3 , <strong>and</strong> so the full set of resultsreadsx 1 = ∆ 1|A| , x 2 = ∆ 2|A| , x 3 = ∆ 3|A| , (8.130)where∣ ∣ b 1 A 12 A 13 ∣∣∣∣∣ A 11 b 1 A 13 ∣∣∣∣∣ ∆ 1 =b 2 A 22 A 23 , ∆ 2 =A 21 b 2 A 23 , ∆ 3 =∣ b 3 A 32 A ∣33 A 31 b 3 A ∣33A 11 A 12 b 1A 21 A 22 b 2A 31 A 32 b 3∣ ∣∣∣∣∣.It can be seen that each Cramer determinant ∆ i is simply |A| but with column ireplaced by the RHS of the original set of equations. If |A| ≠ 0 then (8.130) gives299


MATRICES AND VECTOR SPACESthe unique solution. The proof given here appears to fail if any of the solutionsx i is zero, but it can be shown that result (8.130) is valid even in such a case.◮Use Cramer’s rule to solve the set of simultaneous equations (8.123).Let us again represent these simultaneous equations by the matrix equation Ax = b, i.e.⎛⎝ 2 4 3⎞ ⎛1 −2 −2 ⎠ ⎝ x ⎞ ⎛1x 2⎠ = ⎝ 4 ⎞0 ⎠ .−3 3 2 x 3 −7From (8.58), the determinant of A is given by |A| = 11. Following the discussion givenabove, the three Cramer determinants are4 4 3∆ 1 =0 −2 −2∣ −7 3 2 ∣ , ∆ 2 4 32 =1 0 −2∣ −3 −7 2 ∣ , ∆ 2 4 43 =1 −2 0∣ −3 3 −7 ∣ .These may be evaluated using the properties of determinants listed in subsection 8.9.1<strong>and</strong> we find ∆ 1 = 22, ∆ 2 = −33 <strong>and</strong> ∆ 3 = 44. From (8.130) the solution to the equations(8.123) is given byx 1 = 2211 =2, x 2 = −3311 = −3, x 3 = 4411 =4,which agrees with the solution found in the previous example. ◭At this point it is useful to consider each of the three equations (8.129) as representinga plane in three-dimensional Cartesian coordinates. Using result (7.42)of chapter 7, the sets of components of the vectors normal to the planes are(A 11 ,A 12 ,A 13 ), (A 21 ,A 22 ,A 23 )<strong>and</strong>(A 31 ,A 32 ,A 33 ), <strong>and</strong> using (7.46) the perpendiculardistances of the planes from the origin are given byd i =b i(A2i1+ A 2 i2 + A2 i3) 1/2<strong>for</strong> i =1, 2, 3.Finding the solution(s) to the simultaneous equations above corresponds to findingthe point(s) of intersection of the planes.If there is a unique solution the planes intersect at only a single point. Thishappens if their normals are linearly independent vectors. Since the rows of Arepresent the directions of these normals, this requirement is equivalent to |A| ≠0.If b = (0 0 0) T = 0 then all the planes pass through the origin <strong>and</strong>, since thereis only a single solution to the equations, the origin is that solution.Let us now turn to the cases where |A| = 0. The simplest such case is that inwhich all three planes are parallel; this implies that the normals are all parallel<strong>and</strong> so A is of rank 1. Two possibilities exist:(i) the planes are coincident, i.e. d 1 = d 2 = d 3 , in which case there is aninfinity of solutions;(ii) the planes are not all coincident, i.e. d 1 ≠ d 2 <strong>and</strong>/or d 1 ≠ d 3 <strong>and</strong>/ord 2 ≠ d 3 , in which case there are no solutions.300


8.18 SIMULTANEOUS LINEAR EQUATIONS(a)(b)Figure 8.1 The two possible cases when A is of rank 2. In both cases all thenormals lie in a horizontal plane but in (a) the planes all intersect on a singleline (corresponding to an infinite number of solutions) whilst in (b) there areno common intersection points (no solutions).It is apparent from (8.130) that case (i) occurs when all the Cramer determinantsare zero <strong>and</strong> case (ii) occurs when at least one Cramer determinant is non-zero.The most complicated cases with |A| = 0 are those in which the normals to theplanes themselves lie in a plane but are not parallel. In this case A has rank 2.Again two possibilities exist <strong>and</strong> these are shown in figure 8.1. Just as in therank-1 case, if all the Cramer determinants are zero then we get an infinity ofsolutions (this time on a line). Of course, in the special case in which b = 0 (<strong>and</strong>the system of equations is homogeneous), the planes all pass through the origin<strong>and</strong> so they must intersect on a line through it. If at least one of the Cramerdeterminants is non-zero, we get no solution.These rules may be summarised as follows.(i) |A| ≠0,b ≠ 0: The three planes intersect at a single point that is not theorigin, <strong>and</strong> so there is only one solution, given by both (8.122) <strong>and</strong> (8.130).(ii) |A| ≠0,b = 0: The three planes intersect at the origin only <strong>and</strong> there isonly the trivial solution, x =0.(iii) |A| =0,b ≠ 0, Cramer determinants all zero: There is an infinity ofsolutions either on a line if A is rank 2, i.e. the cofactors are not all zero,or on a plane if A is rank 1, i.e. the cofactors are all zero.(iv) |A| =0,b ≠ 0, Cramer determinants not all zero: No solutions.(v) |A| =0,b = 0: The three planes intersect on a line through the origingiving an infinity of solutions.8.18.3 Singular value decompositionThere exists a very powerful technique <strong>for</strong> dealing with a simultaneous set oflinear equations Ax = b, such as (8.118), which may be applied whether or not301


MATRICES AND VECTOR SPACESthe number of simultaneous equations M is equal to the number of unknowns N.This technique is known as singular value decomposition (SVD) <strong>and</strong> is the methodof choice in analysing any set of simultaneous linear equations.We will consider the general case, in which A is an M × N (complex) matrix.Let us suppose we can write A as the product §A = USV † , (8.131)where the matrices U, S <strong>and</strong> V have the following properties.(i) The square matrix U has dimensions M × M <strong>and</strong> is unitary.(ii) The matrix S has dimensions M × N (thesamedimensionsasthoseofA)<strong>and</strong> is diagonal in the sense that S ij =0ifi ≠ j. We denote its diagonalelements by s i <strong>for</strong> i =1, 2,...,p,wherep = min(M,N); these elements aretermed the singular values of A.(iii) The square matrix V has dimensions N × N <strong>and</strong> is unitary.We must now determine the elements of these matrices in terms of the elements ofA. From the matrix A, we can construct two square matrices: A † A with dimensionsN ×N <strong>and</strong> AA † with dimensions M ×M. Both are clearly Hermitian. From (8.131),<strong>and</strong> using the fact that U <strong>and</strong> V are unitary, we findA † A = VS † U † USV † = VS † SV † (8.132)AA † = USV † VS † U † = USS † U † , (8.133)where S † S <strong>and</strong> SS † are diagonal matrices with dimensions N × N <strong>and</strong> M × Mrespectively. The first p elements of each diagonal matrix are s 2 i , i =1, 2,...,p,where p = min(M,N), <strong>and</strong> the rest (where they exist) are zero.These two equations imply that both V −1 A † AV ( = V −1 A † A(V † ) −1) <strong>and</strong>, bya similar argument, U −1 AA † U, must be diagonal. From our discussion of thediagonalisation of Hermitian matrices in section 8.16, we see that the columns ofV must there<strong>for</strong>e be the normalised eigenvectors v i , i =1, 2,...,N, of the matrixA † A <strong>and</strong> the columns of U must be the normalised eigenvectors u j , j =1, 2,...,M,of the matrix AA † . Moreover, the singular values s i must satisfy s 2 i = λ i ,wherethe λ i are the eigenvalues of the smaller of A † A <strong>and</strong> AA † . Clearly, the λ i arealso some of the eigenvalues of the larger of these two matrices, the remainingones being equal to zero. Since each matrix is Hermitian, the λ i are real <strong>and</strong> thesingular values s i may be taken as real <strong>and</strong> non-negative. Finally, to make thedecomposition (8.131) unique, it is customary to arrange the singular values indecreasing order of their values, so that s 1 ≥ s 2 ≥ ···≥ s p .§ The proof that such a decomposition always exists is beyond the scope of this book. For afull account of SVD one might consult, <strong>for</strong> example, G. H. Golub <strong>and</strong> C. F. Van Loan, MatrixComputations, 3rd edn (Baltimore MD: Johns Hopkins University Press, 1996).302


8.18 SIMULTANEOUS LINEAR EQUATIONS◮Show that, <strong>for</strong> i =1, 2,...,p, Av i = s i u i <strong>and</strong> A † u i = s i v i ,wherep =min(M,N).Post-multiplying both sides of (8.131) by V, <strong>and</strong> using the fact that V is unitary, we obtainAV = US.Since the columns of V <strong>and</strong> U consist of the vectors v i <strong>and</strong> u j respectively <strong>and</strong> S has onlydiagonal non-zero elements, we find immediately that, <strong>for</strong> i =1, 2,...,p,Av i = s i u i . (8.134)Moreover, we note that Av i =0<strong>for</strong>i = p +1,p+2,...,N.Taking the Hermitian conjugate of both sides of (8.131) <strong>and</strong> post-multiplying by U, weobtainA † U = VS † = VS T ,where we have used the fact that U is unitary <strong>and</strong> S is real. We then see immediately that,<strong>for</strong> i =1, 2,...,p,A † u i = s i v i . (8.135)We also note that A † u i =0<strong>for</strong>i = p +1,p+2,...,M. Results (8.134) <strong>and</strong> (8.135) are useful<strong>for</strong> investigating the properties of the SVD. ◭The decomposition (8.131) has some advantageous features <strong>for</strong> the analysis ofsets of simultaneous linear equations. These are best illustrated by writing thedecomposition (8.131) in terms of the vectors u i <strong>and</strong> v i asp∑A = s i u i (v i ) † ,i=1where p = min(M,N). It may be, however, that some of the singular values s iare zero, as a result of degeneracies in the set of M linear equations Ax = b.Let us suppose that there are r non-zero singular values. Since our convention isto arrange the singular values in order of decreasing size, the non-zero singularvalues are s i , i =1, 2,...,r, <strong>and</strong> the zero singular values are s r+1 ,s r+2 ,...,s p .There<strong>for</strong>e we can write A asr∑A = s i u i (v i ) † . (8.136)i=1Let us consider the action of (8.136) on an arbitrary vector x. This is given byr∑Ax = s i u i (v i ) † x.i=1Since (v i ) † x is just a number, we see immediately that the vectors u i , i =1, 2,...,r,must span the range of the matrix A; moreover, these vectors <strong>for</strong>m an orthonormalbasis <strong>for</strong> the range. Further, since this subspace is r-dimensional, we haverank A = r, i.e. the rank of A is equal to the number of non-zero singular values.The SVD is also useful in characterising the null space of A. From (8.119),we already know that the null space must have dimension N − r; so,ifA has r303


MATRICES AND VECTOR SPACESnon-zero singular values s i , i =1, 2,...,r, then from the worked example abovewe haveAv i =0 <strong>for</strong>i = r +1,r+2,...,N.Thus, the N − r vectors v i , i = r +1,r+2,...,N, <strong>for</strong>m an orthonormal basis <strong>for</strong>the null space of A.◮ Find the singular value decompostion of the matrix⎛⎞2 2 2 2⎜ 17 1A = ⎝ − 17 − 1 ⎟10 10 10 10 ⎠ . (8.137)3595− 3 5− 9 5The matrix A has dimension 3 × 4(i.e.M =3,N =4),<strong>and</strong>sowemayconstructfromit the 3 × 3matrixAA † <strong>and</strong> the 4 × 4matrixA † A (in fact, since A is real, the Hermitianconjugates are just transposes). We begin by finding the eigenvalues λ i <strong>and</strong> eigenvectors u iof the smaller matrix AA † . This matrix is easily found to be given by⎛⎞16 0 0AA † 12= ⎝⎠ ,02950125<strong>and</strong> its characteristic equation reads∣ 16 − λ 0 0 ∣∣∣∣∣ 2905∣− λ 125=(16− λ)(36 − 13λ + λ 2 )=0.12 360 − λ 5 5Thus, the eigenvalues are λ 1 = 16, λ 2 =9,λ 3 = 4. Since the singular values of A are givenby s i = √ λ i <strong>and</strong> the matrix S in (8.131) has the same dimensions as A, we have⎛S = ⎝ 4 0 0 0⎞0 3 0 0 ⎠ , (8.138)0 0 2 0where we have arranged the singular values in order of decreasing size. Now the matrix Uhas as its columns the normalised eigenvectors u i of the 3×3 matrixAA † . These normalisedeigenvectors correspond to the eigenvalues of AA † as follows:λ 1 =16 ⇒ u 1 = (1 0 0) T<strong>and</strong> so we obtain the matrix5365λ 2 =9 ⇒ u 2 3=(05λ 3 =4 ⇒ u 3 =(0 − 4 5⎛1 0 0⎜U = ⎝ 03045− 4 5 535⎞45 )T35 )T ,⎟⎠ . (8.139)The columns of the matrix V in (8.131) are the normalised eigenvectors of the 4 × 4matrix A † A, which is given by⎛⎞29 21 3 11A † A = 1 ⎜ 21 29 11 3 ⎟⎝4 3 11 29 21⎠ .11 3 21 29304


8.18 SIMULTANEOUS LINEAR EQUATIONSWe already know from the above discussion, however, that the non-zero eigenvalues ofthis matrix are equal to those of AA † found above, <strong>and</strong> that the remaining eigenvalue iszero. The corresponding normalised eigenvectors are easily found:λ 1 =16 ⇒ v 1 = 1 (1 1 1 1)T2<strong>and</strong>sothematrixV is given byλ 2 =9 ⇒ v 2 = 1 (1 1 − 1 − 1)T2λ 3 =4 ⇒ v 3 = 1 (−1 1 1 − 1)T2λ 4 =0 ⇒ v 4 = 1 (1 − 1 1 − 1)T2⎛V = 1 ⎜⎝2⎞1 1 −1 11 1 1 −1 ⎟1 −1 1 1⎠ . (8.140)1 −1 −1 −1Alternatively, we could have found the first three columns of V by using the relation(8.135) to obtainv i = 1 A † u i <strong>for</strong> i =1, 2, 3.s iThe fourth eigenvector could then be found using the Gram–Schmidt orthogonalisationprocedure. We note that if there were more than one eigenvector corresponding to a zeroeigenvalue then we would need to use this procedure to orthogonalise these eigenvectorsbe<strong>for</strong>e constructing the matrix V.Collecting our results together, we find the SVD of the matrix A:⎛1 0 0A = USV † ⎜= ⎝ 03045− 4 5 535⎛⎞ ⎛⎟⎠ ⎝ 4 0 0 0⎞0 3 0 0 ⎠⎜0 0 2 0 ⎝this can be verified by direct multiplication. ◭1212− 1 212112122− 1 2− 1 21212− 1 212− 1 212− 1 2Let us now consider the use of SVD in solving a set of M simultaneous linearequations in N unknowns, which we write again as Ax = b. Firstly, considerthe solution of a homogeneous set of equations, <strong>for</strong> which b = 0. As mentionedpreviously, if A is square <strong>and</strong> non-singular (<strong>and</strong> so possesses no zero singularvalues) then the equations have the unique trivial solution x = 0. Otherwise,anyof the vectors v i , i = r +1,r+2,...,N, or any linear combination of them, willbe a solution.In the inhomogeneous case, where b is not a zero vector, the set of equationswill possess solutions if b lies in the range of A. To investigate these solutions, itis convenient to introduce the N × M matrix S, which is constructed by takingthe transpose of S in (8.131) <strong>and</strong> replacing each non-zero singular value s i onthe diagonal by 1/s i . It is clear that, with this construction, SS is an M × Mdiagonal matrix with diagonal entries that equal unity <strong>for</strong> those values of j <strong>for</strong>which s j ≠ 0, <strong>and</strong> zero otherwise.Now consider the vectorˆx = VSU † b. (8.141)305⎞⎟⎠ ;


MATRICES AND VECTOR SPACESUsing the unitarity of the matrices U <strong>and</strong> V, we find thatAˆx − b = USSU † b − b = U(SS − I)U † b. (8.142)The matrix (SS − I) is diagonal <strong>and</strong> the jth element on its leading diagonal isnon-zero (<strong>and</strong> equal to −1) only when s j = 0. However, the jth element of thevector U † b is given by the scalar product (u j ) † b;ifb lies in the range of A, thisscalar product can be non-zero only if s j ≠ 0. Thus the RHS of (8.142) mustequal zero, <strong>and</strong> so ˆx given by (8.141) is a solution to the equations Ax = b. Wemay, however, add to this solution any linear combination of the N − r vectors v i ,i = r+1,r+2,...,N, that <strong>for</strong>m an orthonormal basis <strong>for</strong> the null space of A; thus,in general, there exists an infinity of solutions (although it is straight<strong>for</strong>ward toshow that (8.141) is the solution vector of shortest length). The only way in whichthe solution (8.141) can be unique is if the rank r equals N, so that the matrix Adoes not possess a null space; this only occurs if A is square <strong>and</strong> non-singular.If b does not lie in the range of A then the set of equations Ax = b doesnot have a solution. Nevertheless, the vector (8.141) provides the closest possible‘solution’ in a least-squares sense. In other words, although the vector (8.141)does not exactly solve Ax = b, it is the vector that minimises the residualɛ = |Ax − b|,where here the vertical lines denote the absolute value of the quantity theycontain, not the determinant. This is proved as follows.Suppose we were to add some arbitrary vector x ′ to the vector ˆx in (8.141).This would result in the addition of the vector b ′ = Ax ′ to Aˆx − b; b ′ is clearly inthe range of A since any part of x ′ belonging to the null space of A contributesnothing to Ax ′ . We would then have|Aˆx − b + b ′ | = |(USSU † − I)b + b ′ |= |U[(SS − I)U † b + U † b ′ ]|= |(SS − I)U † b + U † b ′ |; (8.143)in the last line we have made use of the fact that the length of a vector is leftunchanged by the action of the unitary matrix U. Now, the jth component of thevector (SS − I)U † b will only be non-zero when s j = 0. However, the jth elementof the vector U † b ′ is given by the scalar product (u j ) † b ′ , which is non-zero only ifs j ≠0,sinceb ′ lies in the range of A. Thus, as these two terms only contribute to(8.143) <strong>for</strong> two disjoint sets of j-values, its minimum value, as x ′ is varied, occurswhen b ′ = 0; this requires x ′ = 0.◮Find the solution(s) to the set of simultaneous linear equations Ax = b, whereA is givenby (8.137) <strong>and</strong> b = (1 0 0) T .To solve the set of equations, we begin by calculating the vector given in (8.141),x = VSU † b,306


8.19 EXERCISESwhere U <strong>and</strong> V are given by (8.139) <strong>and</strong> (8.140) respectively <strong>and</strong> S is obtained by takingthe transpose of S in (8.138) <strong>and</strong> replacing all the non-zero singular values s i by 1/s i . Thus,S reads⎛1⎞0 04 10 03 S = ⎜⎟⎝10 0 ⎠ .20 0 0Substituting the appropriate matrices into the expression <strong>for</strong> x we findx = 1 (1 1 1 8 1)T . (8.144)It is straight<strong>for</strong>ward to show that this solves the set of equations Ax = b exactly, <strong>and</strong>so the vector b = (1 0 0) T must lie in the range of A. This is, in fact, immediatelyclear, since b = u 1 . The solution (8.144) is not, however, unique. There are three non-zerosingular values, but N = 4. Thus, the matrix A has a one-dimensional null space, whichis ‘spanned’ by v 4 , the fourth column of V, given in (8.140). The solutions to our set ofequations, consisting of the sum of the exact solution <strong>and</strong> any vector in the null space ofA, there<strong>for</strong>e lie along the linex = 1 (1 1 1 8 1)T + α(1 − 1 1 − 1) T ,where the parameter α can take any real value. We note that (8.144) is the point on thisline that is closest to the origin. ◭8.19 Exercises8.1 Which of the following statements about linear vector spaces are true? Where astatement is false, give a counter-example to demonstrate this.(a) Non-singular N × N matrices <strong>for</strong>m a vector space of dimension N 2 .(b) Singular N × N matrices <strong>for</strong>m a vector space of dimension N 2 .(c) Complex numbers <strong>for</strong>m a vector space of dimension 2.(d) Polynomial functions of x <strong>for</strong>m an infinite-dimensional vector space.(e) Series {a 0 ,a 1 ,a 2 ,...,a N } <strong>for</strong> which ∑ Nn=0 |a n| 2 =1<strong>for</strong>manN-dimensionalvector space.(f) Absolutely convergent series <strong>for</strong>m an infinite-dimensional vector space.(g) Convergent series with terms of alternating sign <strong>for</strong>m an infinite-dimensionalvector space.8.2 Evaluate the determinants∣ a h g∣∣∣∣∣∣ 1 0 2 3(a)h b f∣ g f c ∣ , (b) 0 1 −2 13 −3 4 −2−2 1 −2 1 ∣<strong>and</strong>gc ge a + ge gb + ge0 b b b(c)c e e b+ e.∣ a b b+ f b+ d ∣307


MATRICES AND VECTOR SPACES8.3 Using the properties of determinants, solve with a minimum of calculation thefollowing equations <strong>for</strong> x:x a a 1a x b 1x +2 x +4 x − 3(a)a b x 1=0, (b)x +3 x x+5∣ a b c 1 ∣∣ x − 2 x − 1 x +1 ∣ =0.8.4 Consider the matrices⎛(a) B =⎝ i 0 −i0 −i i−i i 0⎞⎛ √ √ √ ⎞⎠ , (b) C = √ 1 3 −√ 2 − 3⎝ 1 6 −1 ⎠ .82 0 2Are they (i) real, (ii) diagonal, (iii) symmetric, (iv) antisymmetric, (v) singular,(vi) orthogonal, (vii) Hermitian, (viii) anti-Hermitian, (ix) unitary, (x) normal?8.5 By considering the matricesA =(1 00 0) ( )0 0, B =,3 4show that AB = 0 does not imply that either A or B is the zero matrix, but thatit does imply that at least one of them is singular.8.6 This exercise considers a crystal whose unit cell has base vectors that are notnecessarily mutually orthogonal.(a) The basis vectors of the unit cell of a crystal, with the origin O at one corner,are denoted by e 1 , e 2 , e 3 .ThematrixG has elements G ij ,whereG ij = e i · e j<strong>and</strong> H ij are the elements of the matrix H ≡ G −1 . Show that the vectorsf i = ∑ j H ije j are the reciprocal vectors <strong>and</strong> that H ij = f i · f j .(b) If the vectors u <strong>and</strong> v are given byu = ∑ u i e i , v = ∑ v i f i ,iiobtain expressions <strong>for</strong> |u|, |v|, <strong>and</strong>u · v.(c) If the basis vectors are each of length a <strong>and</strong> the angle between each pair isπ/3, write down G <strong>and</strong> hence obtain H.(d) Calculate (i) the length of the normal from O onto the plane containing thepoints p −1 e 1 , q −1 e 2 , r −1 e 3 , <strong>and</strong> (ii) the angle between this normal <strong>and</strong> e 1 .8.7 Prove the following results involving Hermitian matrices:(a) If A is Hermitian <strong>and</strong> U is unitary then U −1 AU is Hermitian.(b) If A is anti-Hermitian then iA is Hermitian.(c) The product of two Hermitian matrices A <strong>and</strong> B is Hermitian if <strong>and</strong> only ifA <strong>and</strong> B commute.(d) If S is a real antisymmetric matrix then A =(I − S)(I + S) −1 is orthogonal.If A is given by( )cos θ sin θA =− sin θ cos θthen find the matrix S that is needed to express A in the above <strong>for</strong>m.(e) If K is skew-hermitian, i.e. K † = −K, thenV =(I + K)(I − K) −1 is unitary.8.8 A <strong>and</strong> B are real non-zero 3 × 3 matrices <strong>and</strong> satisfy the equation(AB) T + B −1 A = 0.(a) Prove that if B is orthogonal then A is antisymmetric.308


8.19 EXERCISES(b) Without assuming that B is orthogonal, prove that A is singular.8.9 The commutator [X, Y] of two matrices is defined by the equation[X, Y] =XY − YX.Two anticommuting matrices A <strong>and</strong> B satisfyA 2 = I, B 2 = I, [A, B] =2iC.(a) Prove that C 2 = I <strong>and</strong> that [B, C] =2iA.(b) Evaluate [[[A, B], [B, C]], [A, B]].8.10 The four matrices S x , S y , S z <strong>and</strong> I are defined by( ) ( )0 10 −iS x =, S1 0y =,i 0( ) ( )1 01 0S z =, I =,0 −10 1where i 2 = −1. Show that S 2 x = I <strong>and</strong> S x S y = iS z , <strong>and</strong> obtain similar resultsby permutting x, y <strong>and</strong> z. Giventhatv is a vector with Cartesian components(v x ,v y ,v z ), the matrix S(v) is defined asS(v) =v x S x + v y S y + v z S z .Prove that, <strong>for</strong> general non-zero vectors a <strong>and</strong> b,S(a)S(b) =a · b I + i S(a × b).Without further calculation, deduce that S(a) <strong>and</strong>S(b) commute if <strong>and</strong> only if a<strong>and</strong> b are parallel vectors.8.11 A general triangle has angles α, β <strong>and</strong> γ <strong>and</strong> corresponding opposite sides a,b <strong>and</strong> c. Express the length of each side in terms of the lengths of the othertwo sides <strong>and</strong> the relevant cosines, writing the relationships in matrix <strong>and</strong> vector<strong>for</strong>m, using the vectors having components a, b, c <strong>and</strong> cos α, cos β,cos γ. Invert thematrix <strong>and</strong> hence deduce the cosine-law expressions involving α, β <strong>and</strong> γ.8.12 Given a matrixA =⎛⎝ 1 β α 1 00⎞⎠ ,0 0 1where α <strong>and</strong> β are non-zero complex numbers, find its eigenvalues <strong>and</strong> eigenvectors.Find the respective conditions <strong>for</strong> (a) the eigenvalues to be real <strong>and</strong> (b) theeigenvectors to be orthogonal. Show that the conditions are jointly satisfied if<strong>and</strong> only if A is Hermitian.8.13 Using the Gram–Schmidt procedure:(a) construct an orthonormal set of vectors from the following:x 1 = (0 0 1 1) T , x 2 =(1 0 − 1 0) T ,x 3 = (1 2 0 2) T , x 4 =(2 1 1 1) T ;309


MATRICES AND VECTOR SPACES(b) find an orthonormal basis, within a four-dimensional Euclidean space, <strong>for</strong>the subspace spanned by the three vectors (1 2 0 0) T ,(3 − 1 2 0) T<strong>and</strong> (0 0 2 1) T .8.14 If a unitary matrix U is written as A + iB, whereA <strong>and</strong> B are Hermitian withnon-degenerate eigenvalues, show the following:(a) A <strong>and</strong> B commute;(b) A 2 + B 2 = I;(c) The eigenvectors of A are also eigenvectors of B;(d) The eigenvalues of U have unit modulus (as is necessary <strong>for</strong> any unitarymatrix).8.15 Determine which of the matrices below are mutually commuting, <strong>and</strong>, <strong>for</strong> thosethat are, demonstrate that they have a complete set of eigenvectors in common:( ) ( )6 −21 8A =, B =,C =−2 9(−9 −10−10 5), D =8 −11(14 22 118.16 Find the eigenvalues <strong>and</strong> a set of eigenvectors of the matrix⎛⎝ 3 4 −21 3 −1⎞⎠ .−1 −2 2Verify that its eigenvectors are mutually orthogonal.8.17 Find three real orthogonal column matrices, each of which is a simultaneouseigenvector ofA =⎛⎝ 0 0 10 1 01 0 0⎞⎛⎠ <strong>and</strong> B =).⎝ 0 1 11 0 11 1 08.18 Use the results of the first worked example in section 8.14 to evaluate, withoutrepeated matrix multiplication, the expression A 6 x,wherex =(2 4 − 1) T <strong>and</strong>A is the matrix given in the example.8.19 Given that A is a real symmetric matrix with normalised eigenvectors e i ,obtainthe coefficients α i involved when column matrix x, which is the solution ofAx − µx = v, (∗)is exp<strong>and</strong>ed as x = ∑ i α ie i .Hereµ is a given constant <strong>and</strong> v is a given columnmatrix.⎞⎠ .(a) Solve (∗) whenA =⎛⎝ 2 1 1 2 00⎞⎠ ,0 0 3µ =2<strong>and</strong>v = (1 2 3) T .(b) Would (∗) have a solution if µ = 1 <strong>and</strong> (i) v = (1 2 3) T , (ii) v =(2 2 3) T ?310


8.19 EXERCISES8.20 Demonstrate that the matrix⎛⎞A = −6 4 4⎝ 2 0 0⎠3 −1 0is defective, i.e. does not have three linearly independent eigenvectors, by showingthe following:(a) its eigenvalues are degenerate <strong>and</strong>, in fact, all equal;(b) any eigenvector has the <strong>for</strong>m (µ (3µ − 2ν) ν) T .(c) if two pairs of values, µ 1 ,ν 1 <strong>and</strong> µ 2 ,ν 2 , define two independent eigenvectorsv 1 <strong>and</strong> v 2 ,thenany third similarly defined eigenvector v 3 canbewrittenasa linear combination of v 1 <strong>and</strong> v 2 ,i.e.v 3 = av 1 + bv 2 ,wherea = µ 3ν 2 − µ 2 ν 3µ 1 ν 2 − µ 2 ν 1<strong>and</strong> b = µ 1ν 3 − µ 3 ν 1.µ 1 ν 2 − µ 2 ν 1Illustrate (c) using the example (µ 1 ,ν 1 )=(1, 1), (µ 2 ,ν 2 )=(1, 2) <strong>and</strong> (µ 3 ,ν 3 )=(0, 1).Show further that of the <strong>for</strong>many matrix⎛⎝ 6n − 6 4− 2n 4 − 4n2 0 0⎞⎠3 − 3n n− 1 2nis defective, with the same eigenvalues <strong>and</strong> eigenvectors as A.8.21 By finding the eigenvectors of the Hermitian matrix( )10 3iH =,−3i 2construct a unitary matrix U such that U † HU = Λ, where Λ is a real diagonalmatrix.8.22 Use the stationary properties of quadratic <strong>for</strong>ms to determine the maximum <strong>and</strong>minimum values taken by the expressionQ =5x 2 +4y 2 +4z 2 +2xz +2xyon the unit sphere, x 2 + y 2 + z 2 = 1. For what values of x, y <strong>and</strong> z do they occur?8.23 Given that the matrix⎛A = −1 2 −1⎝ 2 −1 0⎞⎠0 −1 2has two eigenvectors of the <strong>for</strong>m (1 y 1) T , use the stationary property of theexpression J(x) =x T Ax/(x T x) to obtain the corresponding eigenvalues. Deducethe third eigenvalue.8.24 Find the lengths of the semi-axes of the ellipse73x 2 +72xy +52y 2 = 100,<strong>and</strong> determine its orientation.8.25 The equation of a particular conic section isQ ≡ 8x 2 1 +8x 2 2 − 6x 1 x 2 = 110.Determine the type of conic section this represents, the orientation of its principalaxes, <strong>and</strong> relevant lengths in the directions of these axes.311


MATRICES AND VECTOR SPACES8.26 Show that the quadratic surface5x 2 +11y 2 +5z 2 − 10yz +2xz − 10xy =4is an ellipsoid with semi-axes of lengths 2, 1 <strong>and</strong> 0.5. Find the direction of itslongest axis.8.27 Find the direction of the axis of symmetry of the quadratic surface7x 2 +7y 2 +7z 2 − 20yz − 20xz +20xy =3.8.28 For the following matrices, find the eigenvalues <strong>and</strong> sufficient of the eigenvectorsto be able to describe the quadratic surfaces associated with them:⎛(a) ⎝ 5 1 −1⎞ ⎛1 5 1 ⎠ , (b) ⎝ 1 2 2⎞ ⎛2 1 2 ⎠ , (c) ⎝ 1 2 1⎞2 4 2 ⎠ .−1 1 52 2 11 2 18.29 This exercise demonstrates the reverse of the usual procedure of diagonalising amatrix.(a) Rearrange the result A ′ = S −1 AS of section 8.16 to express the originalmatrix A in terms of the unitary matrix S <strong>and</strong> the diagonal matrix A ′ . Henceshow how to construct a matrix A that has given eigenvalues <strong>and</strong> given(orthogonal) column matrices as its eigenvectors.(b) Find the matrix that has as eigenvectors (1 2 1) T , (1 − 1 1) T <strong>and</strong>(1 0 − 1) T , with corresponding eigenvalues λ, µ <strong>and</strong> ν.(c) Try a particular case, say λ =3,µ = −2 <strong>and</strong>ν = 1, <strong>and</strong> verify by explicitsolution that the matrix so found does have these eigenvalues.8.30 Find an orthogonal trans<strong>for</strong>mation that takes the quadratic <strong>for</strong>mQ ≡−x 2 1 − 2x 2 2 − x 2 3 +8x 2 x 3 +6x 1 x 3 +8x 1 x 2into the <strong>for</strong>mµ 1 y1 2 + µ 2 y2 2 − 4y3,2<strong>and</strong> determine µ 1 <strong>and</strong> µ 2 (see section 8.17).8.31 One method of determining the nullity (<strong>and</strong> hence the rank) of an M × N matrixA is as follows.• Write down an augmented transpose of A, by adding on the right an N × Nunit matrix <strong>and</strong> thus producing an N × (M + N) array B.• Subtract a suitable multiple of the first row of B from each of the other lowerrows so as to make B i1 =0<strong>for</strong>i>1.• Subtract a suitable multiple of the second row (or the uppermost row thatdoes not start with M zero values) from each of the other lower rows so as tomake B i2 =0<strong>for</strong>i>2.• Continue in this way until all remaining rows have zeros in the first M places.The number of such rows is equal to the nullity of A, <strong>and</strong>theN rightmostentries of these rows are the components of vectors that span the null space.They can be made orthogonal if they are not so already.Use this method to show that the nullity of⎛⎞−1 3 2 73 10 −6 17A = ⎜−1 −2 2 −3⎟⎝ 2 3 −4 4 ⎠4 0 −8 −4312


8.19 EXERCISESis 2 <strong>and</strong> that an orthogonal base <strong>for</strong> the null space of A is provided by any twocolumn matrices of the <strong>for</strong>m (2 + α i − 2α i 1 α i ) T ,<strong>for</strong>whichtheα i (i =1, 2)are real <strong>and</strong> satisfy 6α 1 α 2 +2(α 1 + α 2 )+5=0.8.32 Do the following sets of equations have non-zero solutions? If so, find them.(a) 3x +2y + z =0, x − 3y +2z =0, 2x + y +3z =0.(b) 2x = b(y + z), x =2a(y − z), x =(6a − b)y − (6a + b)z.8.33 Solve the simultaneous equations2x +3y + z =11,x + y + z =6,5x − y +10z =34.8.34 Solve the following simultaneous equations <strong>for</strong> x 1 , x 2 <strong>and</strong> x 3 , using matrixmethods:x 1 +2x 2 +3x 3 =1,3x 1 +4x 2 +5x 3 =2,x 1 +3x 2 +4x 3 =3.8.35 Show that the following equations have solutions only if η = 1 or 2, <strong>and</strong> findthem in these cases:x + y + z =1,x +2y +4z = η,x +4y +10z = η 2 .8.36 Find the condition(s) on α such that the simultaneous equationsx 1 + αx 2 =1,x 1 − x 2 +3x 3 = −1,2x 1 − 2x 2 + αx 3 = −2have (a) exactly one solution, (b) no solutions, or (c) an infinite number ofsolutions; give all solutions where they exist.8.37 Make an LU decomposition of the matrix⎛A = ⎝ 3 6 9⎞1 0 5⎠2 −2 16<strong>and</strong> hence solve Ax = b, where(i)b = (21 9 28) T , (ii) b = (21 7 22) T .8.38 Make an LU decomposition of the matrix⎛⎞2 −3 1 3⎜1 4 −3 −3⎟A = ⎝5 3 −1 −1⎠ .3 −6 −3 1Hence solve Ax = b <strong>for</strong> (i) b =(−4 1 8 −5) T , (ii) b =(−10 0 −3 −24) T .Deduce that det A = −160 <strong>and</strong> confirm this by direct calculation.8.39 Use the Cholesky separation method to determine whether the following matricesare positive definite. For each that is, determine the corresponding lower diagonalmatrix L:⎛A =⎝ 2 1 31 3 −13 −1 1⎞ ⎛⎠ , B =313⎝ 5 0 √ ⎞3√0 3 0 ⎠ .3 0 3


MATRICES AND VECTOR SPACES8.40 Find the equation satisfied by the squares of the singular values of the matrixassociated with the following over-determined set of equations:2x +3y + z =0x − y − z =12x + y =02y + z = −2.Show that one of the singular values is close to zero. Determine the two largersingular values by an appropriate iteration process <strong>and</strong> the smallest one byindirect calculation.8.41 Find the SVD of⎛A = ⎝ 0 −1⎞1 1 ⎠ ,−1 0showing that the singular values are √ 3<strong>and</strong>1.8.42 Find the SVD <strong>for</strong>m of the matrix⎛⎞22 28 −22⎜ 1 −2 −19⎟A = ⎝19 −2 −1⎠ .−6 12 6Use it to determine the best solution x of the equation Ax = b when (i) b =(6 − 39 15 18) T , (ii) b =(9 − 42 15 15) T , showing that (i) has an exactsolution, but that the best solution to (ii) has a residual of √ 18.8.43 Four experimental measurements of particular combinations of three physicalvariables, x, y <strong>and</strong> z, gave the following inconsistent results:13x +22y − 13z =4,10x − 8y − 10z =44,10x − 8y − 10z =47,9x − 18y − 9z =72.Find the SVD best values <strong>for</strong> x, y <strong>and</strong> z. Identify the null space of A <strong>and</strong> henceobtain the general SVD solution.8.20 Hints <strong>and</strong> answers8.1 (a) False. O N ,theN × N null matrix, ( is not ) non-singular. ( )1 0 0 0(b) False. Consider the sum of<strong>and</strong>.0 0 0 1(c) True.(d) True.(e) False. Consider b n = a n + a n <strong>for</strong> which ∑ Nn=0 |b n| 2 =4≠ 1, or note that thereis no zero vector with unit norm.(f) True.(g) False. Consider the two series defined bya 0 = 1 , a 2 n =2(− 1 2 )n <strong>for</strong> n ≥ 1; b n = −(− 1 2 )n <strong>for</strong> n ≥ 0.Theseriesthatisthesumof{a n } <strong>and</strong> {b n } does not have alternating signs<strong>and</strong> so closure does not hold.8.3 (a) x = a, b or c; (b)x = −1; the equation is linear in x.314


8.20 HINTS AND ANSWERS8.5 Use the ( property of the determinant ) of a matrix product.0 − tan(θ/2)8.7 (d) S =.tan(θ/2) 0(e) Note that (I + K)(I − K) =I − K 2 =(I − K)(I + K).8.9 (b) 32iA.8.11 a = b cos γ + c cos β, <strong>and</strong> cyclic permutations; a 2 = b 2 + c 2 − 2bc cos α, <strong>and</strong> cyclicpermutations.8.13 (a) 2 −1/2 (0 0 1 1) T , 6 −1/2 (2 0 − 1 1) T ,39 −1/2 (−1 6 − 1 1) T , 13 −1/2 (2 1 2 − 2) T .(b) 5 −1/2 (1 2 0 0) T , (345) −1/2 (14 − 7 10 0) T ,(18 285) −1/2 (−56 28 98 69) T .8.15 C does not commute with the others; A, B <strong>and</strong> D have (1 − 2) T <strong>and</strong> (2 1) T ascommon eigenvectors.8.17 For A : (1 0 − 1) T , (1 α 1 1) T , (1 α 2 1) T .For B : (1 1 1) T , (β 1 γ 1 − β 1 − γ 1 ) T , (β 2 γ 2 − β 2 − γ 2 ) T .The α i , β i <strong>and</strong> γ i are arbitrary.Simultaneous <strong>and</strong> orthogonal: (1 0 − 1) T , (1 1 1) T , (1 − 2 1) T .8.19 α j =(v · e j∗ )/(λ j − µ), where λ j is the eigenvalue corresponding to e j .(a) x = (2 1 3) T .(b) Since µ is equal to one of A’s eigenvalues λ j , the equation only has a solutionif v · e j∗ = 0; (i) no solution; (ii) x =(1 1 3/2) T .8.21 U = (10) −1/2 (1, 3i;3i, 1), Λ = (1, 0; 0, 11).8.23 J =(2y 2 −4y +4)/(y 2 +2), with stationary values at y = ± √ 2 <strong>and</strong> correspondingeigenvalues 2 ∓ √ 2. From the trace property of A, the third eigenvalue equals 2.8.25 Ellipse; θ = π/4, a = √ 22; θ =3π/4, b = √ 10.8.27 The direction of the eigenvector having the unrepeated eigenvalue is(1, 1, −1)/ √ 3.8.29 (a) A = SA ′ S † ,whereS is the matrix whose columns are the eigenvectors of thematrix A to be constructed, <strong>and</strong> A ′ =diag(λ, µ, ν).(b) A =(λ +2µ +3ν, 2λ − 2µ, λ +2µ − 3ν; 2λ − 2µ, 4λ +2µ, 2λ − 2µ;λ +2µ − 3ν, 2λ − 2µ, λ +2µ +3ν).1(c) (1, 5, −2; 5, 4, 5; −2, 5, 1).38.31 The null space is spanned by (2 0 1 0) T <strong>and</strong> (1 − 2 0 1) T .8.33 x =3,y =1,z =2.8.35 First show that A is singular. η =1,x =1+2z, y = −3z; η =2,x =2z,y =1− 3z.8.37 L =(1, 0, 0; 1 , 1, 0; 2 , 3, 1),3 3U =(3, 6, 9; 0, −2, 2; 0, 0, 4).(i) x =(−1 1 2) T . (ii) x =(−3 2 2) T .8.39 A is not positive definite, as L 33 is calculated to be √ −6.B = LL T , where the non-zero elements of L areL 11 = √ 5,L 31 = √ 3/5, L 22 = √ 3,L 33 = √ 12/5.8.41⎛( )2 1A † A =, U = √ 1 ⎝1 26√ √ ⎞−1 3√ 22 0 2−1 − √ ⎠√, V = 1 (1 1√3 2 2 1 −18.43 The singular values are 12 √ 6, 0, 18 √ 3 <strong>and</strong> the calculated best solution is x =1.71,y = −1.94,z = −1.71. The null space is the line x = z,y = 0 <strong>and</strong> the generalSVD solution is x =1.71 + λ, y = −1.94, z= −1.71 + λ.).315


9Normal modesAny student of the physical sciences will encounter the subject of oscillations onmany occasions <strong>and</strong> in a wide variety of circumstances, <strong>for</strong> example the voltage<strong>and</strong> current oscillations in an electric circuit, the vibrations of a mechanicalstructure <strong>and</strong> the internal motions of molecules. The matrices studied in theprevious chapter provide a particularly simple way to approach what may appear,at first glance, to be difficult physical problems.We will consider only systems <strong>for</strong> which a position-dependent potential exists,i.e., the potential energy of the system in any particular configuration dependsupon the coordinates of the configuration, which need not be be lengths, however;the potential must not depend upon the time derivatives (generalised velocities) ofthese coordinates. So, <strong>for</strong> example, the potential −qv · A used in the Lagrangi<strong>and</strong>escription of a charged particle in an electromagnetic field is excluded. Afurther restriction that we place is that the potential has a local minimum atthe equilibrium point; physically, this is a necessary <strong>and</strong> sufficient condition <strong>for</strong>stable equilibrium. By suitably defining the origin of the potential, we may takeits value at the equilibrium point as zero.We denote the coordinates chosen to describe a configuration of the systemby q i , i =1, 2,...,N.Theq i need not be distances; some could be angles, <strong>for</strong>example. For convenience we can define the q i so that they are all zero at theequilibrium point. The instantaneous velocities of various parts of the system willdepend upon the time derivatives of the q i , denoted by ˙q i . For small oscillationsthe velocities will be linear in the ˙q i <strong>and</strong> consequently the total kinetic energy Twill be quadratic in them – <strong>and</strong> will include cross terms of the <strong>for</strong>m ˙q i˙q j withi ≠ j. The general expression <strong>for</strong> T can be written as the quadratic <strong>for</strong>mT = ∑ ∑a ij ˙q i ˙q j = ˙q T A˙q, (9.1)i jwhere ˙q is the column vector (˙q 1 ˙q 2 ··· ˙q N ) T <strong>and</strong> the N × N matrix Ais real <strong>and</strong> may be chosen to be symmetric. Furthermore, A, like any matrix316


9.1 TYPICAL OSCILLATORY SYSTEMScorresponding to a kinetic energy, is positive definite; that is, whatever non-zeroreal values the ˙q i take, the quadratic <strong>for</strong>m (9.1) has a value > 0.Turning now to the potential energy, we may write its value <strong>for</strong> a configurationq by means of a Taylor expansion about the origin q = 0,V (q) =V (0)+ ∑ i∂V(0)∂q iq i + 1 2∑ ∑ij∂ 2 V (0)∂q i ∂q jq i q j + ··· .However, we have chosen V (0) = 0 <strong>and</strong>, since the origin is an equilibrium point,there is no <strong>for</strong>ce there <strong>and</strong> ∂V(0)/∂q i = 0. Consequently, to second order in theq i we also have a quadratic <strong>for</strong>m, but in the coordinates rather than in their timederivatives:V = ∑ ∑b ij q i q j = q T Bq, (9.2)i jwhere B is, or can be made, symmetric. In this case, <strong>and</strong> in general, the requirementthat the potential is a minimum means that the potential matrix B, like the kineticenergy matrix A, is real <strong>and</strong> positive definite.9.1 Typical oscillatory systemsWe now introduce particular examples, although the results of this section aregeneral, given the above restrictions, <strong>and</strong> the reader will find it easy to apply theresults to many other instances.Consider first a uni<strong>for</strong>m rod of mass M <strong>and</strong> length l, attached by a light stringalso of length l to a fixed point P <strong>and</strong> executing small oscillations in a verticalplane. We choose as coordinates the angles θ 1 <strong>and</strong> θ 2 shown, with exaggeratedmagnitude, in figure 9.1. In terms of these coordinates the centre of gravity of therod has, to first order in the θ i , a velocity component in the x-direction equal tol˙θ 1 + 1 2 l˙θ 2 <strong>and</strong>inthey-direction equal to zero. Adding in the rotational kineticenergy of the rod about its centre of gravity we obtain, to second order in the ˙θ i ,T ≈ 1 2 Ml2 (˙θ 2 1 + 1 4 ˙θ 2 2 + ˙θ 1˙θ2 )+ 1 24 Ml2˙θ2 2= 1 6 Ml2 ( 3˙θ 2 1 +3˙θ 1˙θ2 + ˙θ 2 2)=112 Ml2˙q T ( 6 33 2)˙q, (9.3)where ˙q T =(˙θ 1˙θ2 ). The potential energy is given byso thatV = Mlg [ (1 − cos θ 1 )+ 1 2 (1 − cos θ 2) ] (9.4)( ) 6 0V ≈ 1 4 Mlg(2θ2 1 + θ2)= 2 112 MlgqT q, (9.5)0 3where g is the acceleration due to gravity <strong>and</strong> q =(θ 1second order in the θ i .317θ 2 ) T ; (9.5) is valid to


NORMAL MODESP P Pθ 1θ 2θ 2lθ 1θ 1θ 2(a) (b) (c)lFigure 9.1 A uni<strong>for</strong>m rod of length l attached to the fixed point P by a lightstring of the same length: (a) the general coordinate system; (b) approximationto the normal mode with lower frequency; (c) approximation to the mode withhigher frequency.With these expressions <strong>for</strong> T <strong>and</strong> V we now apply the conservation of energy,d(T + V )=0, (9.6)dtassuming that there are no external <strong>for</strong>ces other than gravity. In matrix <strong>for</strong>m(9.6) becomesddt (˙qT A˙q + q T Bq) =¨q T A˙q + ˙q T A¨q + ˙q T Bq + q T B˙q =0,which, using A = A T <strong>and</strong> B = B T , gives2˙q T (A¨q + Bq) =0.We will assume, although it is not clear that this gives the only possible solution,that the above equation implies that the coefficient of each ˙q i is separately zero.HenceA¨q + Bq = 0. (9.7)For a rigorous derivation Lagrange’s equations should be used, as in chapter 22.Nowwesearch<strong>for</strong>setsofcoordinatesq that all oscillate with the same period,i.e. the total motion repeats itself exactly after a finite interval. Solutions of this<strong>for</strong>m will satisfyq = x cos ωt; (9.8)the relative values of the elements of x in such a solution will indicate how each318


9.1 TYPICAL OSCILLATORY SYSTEMScoordinate is involved in this special motion. In general there will be N valuesof ω if the matrices A <strong>and</strong> B are N × N <strong>and</strong> these values are known as normalfrequencies or eigenfrequencies.Putting (9.8) into (9.7) yields−ω 2 Ax + Bx =(B − ω 2 A)x = 0. (9.9)Our work in section 8.18 showed that this can have non-trivial solutions only if|B − ω 2 A| =0. (9.10)This is a <strong>for</strong>m of characteristic equation <strong>for</strong> B, except that the unit matrix I hasbeen replaced by A. It has the more familiar <strong>for</strong>m if a choice of coordinates ismade in which the kinetic energy T is a simple sum of squared terms, i.e. it hasbeen diagonalised, <strong>and</strong> the scale of the new coordinates is then chosen to makeeach diagonal element unity.However, even in the present case, (9.10) can be solved to yield ω 2 k <strong>for</strong> k =1, 2,...,N,whereN is the order of A <strong>and</strong> B. The values of ω k can be usedwith (9.9) to find the corresponding column vector x k <strong>and</strong> the initial (stationary)physical configuration that, on release, will execute motion with period 2π/ω k .In equation (8.76) we showed that the eigenvectors of a real symmetric matrixwere, except in the case of degeneracy of the eigenvalues, mutually orthogonal.In the present situation an analogous, but not identical, result holds. It is shownin section 9.3 that if x 1 <strong>and</strong> x 2 are two eigenvectors satisfying (9.9) <strong>for</strong> differentvalues of ω 2 then they are orthogonal in the sense that(x 2 ) T Ax 1 =0 <strong>and</strong> (x 2 ) T Bx 1 =0.The direct ‘scalar product’ (x 2 ) T x 1 , <strong>for</strong>mally equal to (x 2 ) T Ix 1 ,isnot,ingeneral,equal to zero.Returning to the suspended rod, we find from (9.10)∣)− ω2 Ml 2( 6 00 3( 6 33 2)∣ Mlg∣∣∣∣=0.1212Writing ω 2 l/g = λ, this becomes∣ 6 − 6λ −3λ−3λ 3 − 2λ ∣ =0 ⇒ λ2 − 10λ +6=0,which has roots λ =5± √ 19. Thus we find that the two normal frequencies aregiven by ω 1 =(0.641g/l) 1/2 <strong>and</strong> ω 2 =(9.359g/l) 1/2 . Putting the lower of the twovalues <strong>for</strong> ω 2 , namely (5 − √ 19)g/l, into (9.9) shows that <strong>for</strong> this modex 1 : x 2 =3(5− √ 19) : 6( √ 19 − 4) = 1.923 : 2.153.This corresponds to the case where the rod <strong>and</strong> string are almost straight out, i.e.they almost <strong>for</strong>m a simple pendulum. Similarly it may be shown that the higher319


NORMAL MODESfrequency corresponds to a solution where the string <strong>and</strong> rod are moving withopposite phase <strong>and</strong> x 1 : x 2 =9.359 : −16.718. The two situations are shown infigure 9.1.In connection with quadratic <strong>for</strong>ms it was shown in section 8.17 how to makea change of coordinates such that the matrix <strong>for</strong> a particular <strong>for</strong>m becomesdiagonal. In exercise 9.6 a method is developed <strong>for</strong> diagonalising simultaneouslytwo quadratic <strong>for</strong>ms (though the trans<strong>for</strong>mation matrix may not be orthogonal).If this process is carried out <strong>for</strong> A <strong>and</strong> B in a general system undergoing stableoscillations, the kinetic <strong>and</strong> potential energies in the new variables η i take the<strong>for</strong>msT = ∑ iV = ∑ iµ i˙η 2 i = ˙η T M˙η, M = diag (µ 1 ,µ 2 ,...,µ N ), (9.11)ν i η 2 i = η T Nη, N = diag (ν 1 ,ν 2 ...,ν N ), (9.12)<strong>and</strong> the equations of motion are the uncoupled equationsµ i¨η i + ν i η i =0, i =1, 2,...,N. (9.13)Clearly a simple renormalisation of the η i can be made that reduces all the µ iin (9.11) to unity. When this is done the variables so <strong>for</strong>med are called normalcoordinates <strong>and</strong> equations (9.13) the normal equations.When a system is executing one of these simple harmonic motions it is said tobe in a normal mode, <strong>and</strong> once started in such a mode it will repeat its motionexactly after each interval of 2π/ω i . Any arbitrary motion of the system maybe written as a superposition of the normal modes, <strong>and</strong> each component modewill execute harmonic motion with the corresponding eigenfrequency; however,unless by chance the eigenfrequencies are in integer relationship, the system willnever return to its initial configuration after any finite time interval.As a second example we will consider a number of masses coupled together bysprings. For this type of situation the potential <strong>and</strong> kinetic energies are automaticallyquadratic functions of the coordinates <strong>and</strong> their derivatives, provided theelastic limits of the springs are not exceeded, <strong>and</strong> the oscillations do not have tobe vanishingly small <strong>for</strong> the analysis to be valid.◮Find the normal frequencies <strong>and</strong> modes of oscillation of three particles of masses m, µm,m connected in that order in a straight line by two equal light springs of <strong>for</strong>ce constant k.This arrangement could serve as a model <strong>for</strong> some linear molecules, e.g. CO 2 .The situation is shown in figure 9.2; the coordinates of the particles, x 1 , x 2 , x 3 , aremeasured from their equilibrium positions, at which the springs are neither extended norcompressed.The kinetic energy of the system is simplyT = 1 m ( ẋ 2 2 1 + µ ẋ 2 2 + ẋ3) 2 ,320


9.1 TYPICAL OSCILLATORY SYSTEMSmk µmkmx 1 x 2 x 3Figure 9.2 Three masses m, µm <strong>and</strong> m connected by two equal light springsof <strong>for</strong>ce constant k.(a)(b)(c)Figure 9.3 The normal modes of the masses <strong>and</strong> springs of a linear moleculesuch as CO 2 .(a)ω 2 =0;(b)ω 2 = k/m; (c)ω 2 =[(µ +2)/µ](k/m).whilst the potential energy stored in the springs isV = 1 k [ (x2 2 − x 1 ) 2 +(x 3 − x 2 ) 2] .The kinetic- <strong>and</strong> potential-energy symmetric matrices are thus⎛A = m ⎝ 1 0 0⎞⎛0 µ 0 ⎠ , B = k ⎝ 1 −1 0−1 2 −12 0 0 12 0 −1 1From (9.10), to find the normal frequencies we have to solve |B − ω 2 A| =0. Thus, writingmω 2 /k = λ, we have1 − λ −1 0−1 2− µλ −1∣ 0 −1 1− λ ∣ =0,which leads to λ =0,1or1+2/µ. The corresponding eigenvectors are respectively⎛x 1 = √ 1 ⎝ 1 ⎞⎛1 ⎠ , x 2 = √ 1 ⎝ 1 ⎞⎛0 ⎠ , x 3 1= √ ⎝ 1⎞−2/µ ⎠ .3 12 −12+(4/µ2 ) 1The physical motions associated with these normal modes are illustrated in figure 9.3.The first, with λ = ω =0<strong>and</strong>allthex i equal, merely describes bodily translation of thewhole system, with no (i.e. zero-frequency) internal oscillations.In the second solution the central particle remains stationary, x 2 = 0, whilst the othertwo oscillate with equal amplitudes in antiphase with each other. This motion, which hasfrequency ω =(k/m) 1/2 , is illustrated in figure 9.3(b).321⎞⎠ .


NORMAL MODESThe final <strong>and</strong> most complicated of the three normal modes has angular frequencyω = {[(µ +2)/µ](k/m)} 1/2 , <strong>and</strong> involves a motion of the central particle which is inantiphase with that of the two outer ones <strong>and</strong> which has an amplitude 2/µ timesasgreat.In this motion (see figure 9.3(c)) the two springs are compressed <strong>and</strong> extended in turn. Wealso note that in the second <strong>and</strong> third normal modes the centre of mass of the moleculeremains stationary. ◭9.2 Symmetry <strong>and</strong> normal modesIt will have been noticed that the system in the above example has an obvioussymmetry under the interchange of coordinates 1 <strong>and</strong> 3: the matrices A <strong>and</strong> B,the equations of motion <strong>and</strong> the normal modes illustrated in figure 9.3 are allunaltered by the interchange of x 1 <strong>and</strong> −x 3 . This reflects the more general resultthat <strong>for</strong> each physical symmetry possessed by a system, there is at least onenormal mode with the same symmetry.The general question of the relationship between the symmetries possessed bya physical system <strong>and</strong> those of its normal modes will be taken up more <strong>for</strong>mallyin chapter 29 where the representation theory of groups is considered. However,we can show here how an appreciation of a system’s symmetry properties willsometimes allow its normal modes to be guessed (<strong>and</strong> then verified), somethingthat is particularly helpful if the number of coordinates involved is greater thantwo <strong>and</strong> the corresponding eigenvalue equation (9.10) is a cubic or higher-degreepolynomial equation.Consider the problem of determining the normal modes of a system consistingof four equal masses M at the corners of a square of side 2L, each pairof masses being connected by a light spring of modulus k that is unstretchedin the equilibrium situation. As shown in figure 9.4, we introduce Cartesiancoordinates x n ,y n , with n =1, 2, 3, 4, <strong>for</strong> the positions of the masses <strong>and</strong> denotetheir displacements from their equilibrium positions R n by q n = x n i + y n j.Thusr n = R n + q n with R n = ±Li ± Lj.The coordinates <strong>for</strong> the system are thus x 1 ,y 1 ,x 2 ,...,y 4 <strong>and</strong> the kinetic energymatrix A is given trivially by MI 8 , where I 8 is the 8 × 8 identity matrix.The potential energy matrix B is much more difficult to calculate <strong>and</strong> involves,<strong>for</strong> each pair of values m, n, evaluating the quadratic approximation to theexpressionb mn = 1 2 k ( |r m − r n |−|R m − R n | ) 2.Expressing each r i in terms of q i <strong>and</strong> R i <strong>and</strong> making the normal assumption that322


9.2 SYMMETRY AND NORMAL MODESy 1 y 2x 2MkMx 1kkkkMx 3ky 3 y 4x 4MFigure 9.4 The arrangement of four equal masses <strong>and</strong> six equal springsdiscussed in the text. The coordinate systems x n ,y n <strong>for</strong> n =1, 2, 3, 4 measurethe displacements of the masses from their equilibrium positions.|R m − R n |≫|q m − q n |,weobtainb mn (= b nm ):b mn = 1 2 k [ |(R m − R n )+(q m − q n )|−|R m − R n | ] 2= 1 2 k { [|Rm− R n | 2 +2(q m − q n ) · (R M − R n )+|q m − q n )| 2] 1/2−|Rm − R n |{ [= 1 2 k|R m − R n | 2 1+ 2(q ] 1/2 2m − q n ) · (R M − R n )|R m − R n | 2 + ··· − 1}{ }≈ 1 2 k (qm − q n ) · (R M − R n ) 2.|R m − R n |This final expression is readily interpretable as the potential energy stored in thespring when it is extended by an amount equal to the component, along theequilibrium direction of the spring, of the relative displacement of its two ends.Applying this result to each spring in turn gives the following expressions <strong>for</strong>the elements of the potential matrix.m n 2b mn /k1 2 (x 1 − x 2 ) 21 3 (y 1 − y 3 ) 21 412 (−x 1 + x 4 + y 1 − y 4 ) 22 312 (x 2 − x 3 + y 2 − y 3 ) 22 4 (y 2 − y 4 ) 23 4 (x 3 − x 4 ) 2 .} 2323


NORMAL MODESThe potential matrix is thus constructed as⎛⎞3 −1 −2 0 0 0 −1 1−1 3 0 0 0 −2 1 −1−2 0 3 1 −1 −1 0 0B = k 0 0 1 3 −1 −1 0 −2.40 0 −1 −1 3 1 −2 0⎜0 −2 −1 −1 1 3 0 0⎟⎝ −1 1 0 0 −2 0 3 −1 ⎠1 −1 0 −2 0 0 −1 3To solve the eigenvalue equation |B − λA| = 0 directly would mean solvingan eigth-degree polynomial equation. Fortunately, we can exploit intuition <strong>and</strong>the symmetries of the system to obtain the eigenvectors <strong>and</strong> correspondingeigenvalues without such labour.Firstly, we know that bodily translation of the whole system, without anyinternal vibration, must be possible <strong>and</strong> that there will be two independentsolutions of this <strong>for</strong>m, corresponding to translations in the x- <strong>and</strong>y- directions.The eigenvector <strong>for</strong> the first of these (written in row <strong>for</strong>m to save space) isEvaluation of Bx (1) givesx (1) = (1 0 1 0 1 0 1 0) T .Bx (1) = (0 0 0 0 0 0 0 0) T ,showing that x (1) is a solution of (B − ω 2 A)x = 0 corresponding to the eigenvalueω 2 = 0, whatever <strong>for</strong>m Ax may take. Similarly,x (2) = (0 1 0 1 0 1 0 1) Tis a second eigenvector corresponding to the eigenvalue ω 2 =0.The next intuitive solution, again involving no internal vibrations, <strong>and</strong>, there<strong>for</strong>e,expected to correspond to ω 2 = 0, is pure rotation of the whole systemabout its centre. In this mode each mass moves perpendicularly to the line joiningits position to the centre, <strong>and</strong> so the relevant eigenvector isx (3) = 1 √2(1 1 1 − 1 − 1 1 − 1 − 1) T .It is easily verified that Bx (3) = 0 thus confirming both the eigenvector <strong>and</strong> thecorresponding eigenvalue. The three non-oscillatory normal modes are illustratedin diagrams (a)–(c) of figure 9.5.We now come to solutions that do involve real internal oscillations, <strong>and</strong>,because of the four-fold symmetry of the system, we expect one of them to be amode in which all the masses move along radial lines – the so-called ‘breathing324


9.2 SYMMETRY AND NORMAL MODES(a) ω 2 =0 (b)ω 2 =0 (c)ω 2 =0 (d) ω 2 =2k/M(e) ω 2 = k/M (f) ω 2 = k/M (g) ω 2 = k/M (h) ω 2 = k/MFigure 9.5 The displacements <strong>and</strong> frequencies of the eight normal modes ofthe system shown in figure 9.4. Modes (a), (b) <strong>and</strong> (c) are not true oscillations:(a) <strong>and</strong> (b) are purely translational whilst (c) is a mode of bodily rotation.Mode (d), the ‘breathing mode’, has the highest frequency <strong>and</strong> the remainingfour, (e)–(h), of lower frequency, are degenerate.mode’. Expressing this motion in coordinate <strong>for</strong>m gives as the fourth eigenvectorx (4) = 1 √2(−1 1 1 1 − 1 − 1 1 − 1) T .Evaluation of Bx (4) yieldsBx (4) =k4 √ 2 (−8 8 8 8 − 8 − 8 8 − 8)T =2kx (4) ,i.e. a multiple of x (4) , confirming that it is indeed an eigenvector. Further, sinceAx (4) = Mx (4) , it follows from (B − ω 2 A)x = 0 that ω 2 =2k/M <strong>for</strong>thisnormalmode. Diagram (d) of the figure illustrates the corresponding motions of the fourmasses.As the next step in exploiting the symmetry properties of the system we notethat, because of its reflection symmetry in the x-axis, the system is invariant underthe double interchange of y 1 with −y 3 <strong>and</strong> y 2 with −y 4 . This leads us to try aneigenvector of the <strong>for</strong>mx (5) =(0 α 0 β 0 − α 0 − β) T .Substituting this trial vector into (B − ω 2 A)x = 0 gives, of course, eight simulta-325


NORMAL MODESneous equations <strong>for</strong> α <strong>and</strong> β, but they are all equivalent to just two, namelyα + β =0,5α + β = 4Mω2 α;kthese have the solution α = −β <strong>and</strong> ω 2 = k/M. The latter thus gives the frequencyof the mode with eigenvectorx (5) =(0 1 0 − 1 0 − 1 0 1) T .Note that, in this mode, when the spring joining masses 1 <strong>and</strong> 3 is most stretched,the one joining masses 2 <strong>and</strong> 4 is at its most compressed. Similarly, based onreflection symmetry in the y-axis,x (6) =(1 0 − 1 0 − 1 0 1 0) Tcan be shown to be an eigenvector corresponding to the same frequency. Thesetwo modes are shown in diagrams (e) <strong>and</strong> (f) of figure 9.5.This accounts <strong>for</strong> six of the expected eight modes, <strong>and</strong> the other two could befound by considering motions that are symmetric about both diagonals of thesquare or are invariant under successive reflections in the x- <strong>and</strong>y- axes. However,since A is a multiple of the unit matrix, <strong>and</strong> since we know that (x (j) ) T Ax (i) =0ifi ≠ j, we can find the two remaining eigenvectors more easily by requiring themto be orthogonal to each of those found so far.Let us take the next (seventh) eigenvector, x (7) , to be given byx (7) =(a b c d e f g h) T .Then orthogonality with each of the x (n) <strong>for</strong> n =1, 2,...,6 yields six equationssatisfied by the unknowns a,b,...,h. As the reader may verify, they can be reducedto the six simple equationsa + g =0, d+ f =0, a+ f = d + g,b + h =0, c+ e =0, b+ c = e + h.With six homogeneous equations <strong>for</strong> eight unknowns, effectively separated intotwo groups of four, we may pick one in each group arbitrarily. Taking a = b =1gives d = e = 1 <strong>and</strong> c = f = g = h = −1 as a solution. Substitution ofx (7) =(1 1 − 1 1 1 − 1 − 1 − 1) T .into the eigenvalue equation checks that it is an eigenvector <strong>and</strong> shows that thecorresponding eigenfrequency is given by ω 2 = k/M.We now have the eigenvectors <strong>for</strong> seven of the eight normal modes <strong>and</strong> theeighth can be found by making it simultaneously orthogonal to each of the otherseven. It is left to the reader to show (or verify) that the final solution isx (8) =(1 − 1 1 1 − 1 − 1 − 1 1) T326


9.3 RAYLEIGH–RITZ METHOD<strong>and</strong> that this mode has the same frequency as three of the other modes. Thegeneral topic of the degeneracy of normal modes is discussed in chapter 29. Themovements associated with the final two modes are shown in diagrams (g) <strong>and</strong>(h) of figure 9.5; this figure summarises all eight normal modes <strong>and</strong> frequencies.Although this example has been lengthy to write out, we have seen that theactual calculations are quite simple <strong>and</strong> provide the full solution to what is<strong>for</strong>mally a matrix eigenvalue equation involving 8 × 8 matrices. It should benoted that our exploitation of the intrinsic symmetries of the system played acrucial part in finding the correct eigenvectors <strong>for</strong> the various normal modes.9.3 Rayleigh–Ritz methodWe conclude this chapter with a discussion of the Rayleigh–Ritz method <strong>for</strong>estimating the eigenfrequencies of an oscillating system. We recall from theintroduction to the chapter that <strong>for</strong> a system undergoing small oscillations thepotential <strong>and</strong> kinetic energy are given byV = q T Bq <strong>and</strong> T = ˙q T A˙q,where the components of q are the coordinates chosen to represent the configurationof the system <strong>and</strong> A <strong>and</strong> B are symmetric matrices (or may be chosen to besuch). We also recall from (9.9) that the normal modes x i <strong>and</strong> the eigenfrequenciesω i are given by(B − ω 2 i A)x i = 0. (9.14)It may be shown that the eigenvectors x i corresponding to different normal modesare linearly independent <strong>and</strong> so <strong>for</strong>m a complete set. Thus, any coordinate vectorq can be written q = ∑ j c jx j . We now consider the value of the generalisedquadratic <strong>for</strong>m∑λ(x) = xT Bxx T Ax = ∑ m(xm ) T c ∗ mB ∑ i c ix ij (xj ) T c ∗ j A ∑ k c kx k ,which, since both numerator <strong>and</strong> denominator are positive definite, is itself nonnegative.Equation (9.14) can be used to replace Bx i , with the result thatλ(x) ==∑m (xm ) T c ∗ mA ∑ ∑i ω2 i c ix ij (xj ) T c ∗ j A ∑ k c kx k∑m (xm ) T c ∗ m∑i ω2 i c iAx i∑j (xj ) T c ∗ j A ∑ k c kx k . (9.15)Now the eigenvectors x i obtained by solving (B − ω 2 A)x = 0 are not mutuallyorthogonal unless either A or B is a multiple of the unit matrix. However, it may327


NORMAL MODESbe shown that they do possess the desirable properties(x j ) T Ax i =0 <strong>and</strong> (x j ) T Bx i =0 ifi ≠ j. (9.16)This result is proved as follows. From (9.14) it is clear that, <strong>for</strong> general i <strong>and</strong> j,(x j ) T (B − ω 2 i A)x i = 0. (9.17)But, by taking the transpose of (9.14) with i replaced by j <strong>and</strong> recalling that A<strong>and</strong> B are real <strong>and</strong> symmetric, we obtain(x j ) T (B − ω 2 j A) =0.Forming the scalar product of this with x i <strong>and</strong> subtracting the result from (9.17)gives(ω 2 j − ω 2 i )(x j ) T Ax i =0.Thus, <strong>for</strong> i ≠ j <strong>and</strong> non-degenerate eigenvalues ωi2 <strong>and</strong> ωj 2 , we have that(x j ) T Ax i = 0, <strong>and</strong> substituting this into (9.17) immediately establishes the correspondingresult <strong>for</strong> (x j ) T Bx i . Clearly, if either A or B is a multiple of the unitmatrix then the eigenvectors are mutually orthogonal in the normal sense. Theorthogonality relations (9.16) are derived again, <strong>and</strong> extended, in exercise 9.6.Using the first of the relationships (9.16) to simplify (9.15), we find thatλ(x) =∑i |c i| 2 ω 2 i (xi ) T Ax i∑k |c k| 2 (x k ) T Ax k . (9.18)Now, if ω0 2 is the lowest eigenfrequency then ω2 i ≥ ω0 2 <strong>for</strong> all i <strong>and</strong>, further, since(x i ) T Ax i ≥ 0 <strong>for</strong> all i the numerator of (9.18) is ≥ ω02 ∑i |c i| 2 (x i ) T Ax i .Henceλ(x) ≡ xT Bxx T Ax ≥ ω2 0, (9.19)<strong>for</strong> any x whatsoever (whether x is an eigenvector or not). Thus we are able toestimate the lowest eigenfrequency of the system by evaluating λ <strong>for</strong> a varietyof vectors x, the components of which, it will be recalled, give the ratios of thecoordinate amplitudes. This is sometimes a useful approach if many coordinatesare involved <strong>and</strong> direct solution <strong>for</strong> the eigenvalues is not possible.An additional result is that the maximum eigenfrequency ωm2 may also beestimated. It is obvious that if we replace the statement ‘ωi2 ≥ ω0 2 <strong>for</strong> all i’ by‘ωi 2 ≤ ωm 2 <strong>for</strong> all i’, then λ(x) ≤ ω2 m <strong>for</strong> any x. Thus λ(x) always lies betweenthe lowest <strong>and</strong> highest eigenfrequencies of the system. Furthermore, λ(x) has astationary value, equal to ωk 2 ,whenx is the kth eigenvector (see subsection 8.17.1).328


9.4 EXERCISES◮Estimate the eigenfrequencies of the oscillating rod of section 9.1.Firstly we recall thatA = Ml212(6 3)3 2<strong>and</strong>B = Mlg12( )6 0.0 3Physical intuition suggests that the slower mode will have a configuration approximatingthat of a simple pendulum (figure 9.1), in which θ 1 = θ 2 , <strong>and</strong> so we use this as a trialvector. Takingx =(θ θ) T ,λ(x) = xT Bxx T Ax = 3Mlgθ2 /47Ml 2 θ 2 /6 = 9g14l =0.643 g l ,<strong>and</strong> we conclude from (9.19) that the lower (angular) frequency is ≤ (0.643g/l) 1/2 .Wehave already seen on p. 319 that the true answer is (0.641g/l) 1/2 <strong>and</strong> so we have comevery close to it.Next we turn to the higher frequency. Here, a typical pattern of oscillation is not soobvious but, rather preempting the answer, we try θ 2 = −2θ 1 ; we then obtain λ =9g/l<strong>and</strong> so conclude that the higher eigenfrequency ≥ (9g/l) 1/2 . We have already seen that theexact answer is (9.359g/l) 1/2 <strong>and</strong> so again we have come close to it. ◭A simplified version of the Rayleigh–Ritz method may be used to estimate theeigenvalues of a symmetric (or in general Hermitian) matrix B, the eigenvectorsof which will be mutually orthogonal. By repeating the calculations leading to(9.18), A being replaced by the unit matrix I, it is easily verified that ifλ(x) = xT Bxx T xis evaluated <strong>for</strong> any vector x thenλ 1 ≤ λ(x) ≤ λ m ,where λ 1 ,λ 2 ...,λ m are the eigenvalues of B in order of increasing size. A similarresult holds <strong>for</strong> Hermitian matrices.9.4 Exercises9.1 Three coupled pendulums swing perpendicularly to the horizontal line containingtheir points of suspension, <strong>and</strong> the following equations of motion are satisfied:−mẍ 1 = cmx 1 + d(x 1 − x 2 ),−Mẍ 2 = cMx 2 + d(x 2 − x 1 )+d(x 2 − x 3 ),−mẍ 3 = cmx 3 + d(x 3 − x 2 ),where x 1 , x 2 <strong>and</strong> x 3 are measured from the equilibrium points; m, M <strong>and</strong> mare the masses of the pendulum bobs; <strong>and</strong> c <strong>and</strong> d are positive constants. Findthe normal frequencies of the system <strong>and</strong> sketch the corresponding patterns ofoscillation. What happens as d → 0ord →∞?9.2 A double pendulum, smoothly pivoted at A, consists of two light rigid rods, AB<strong>and</strong> BC, each of length l, which are smoothly jointed at B <strong>and</strong> carry masses m <strong>and</strong>αm at B <strong>and</strong> C respectively. The pendulum makes small oscillations in one plane329


NORMAL MODESunder gravity. At time t, AB <strong>and</strong> BC make angles θ(t) <strong>and</strong>φ(t), respectively, withthe downward vertical. Find quadratic expressions <strong>for</strong> the kinetic <strong>and</strong> potentialenergies of the system <strong>and</strong> hence show that the normal modes have angularfrequencies given byω 2 = g [1+α ± √ ]α(1 + α) .lFor α =1/3, show that in one of the normal modes the mid-point of BC doesnot move during the motion.9.3 Continue the worked example, modelling a linear molecule, discussed at the endofsection9.1,<strong>for</strong>thecaseinwhichµ =2.(a) Show that the eigenvectors derived there have the expected orthogonalityproperties with respect to both A <strong>and</strong> B.(b) For the situation in which the atoms are released from rest with initialdisplacements x 1 =2ɛ, x 2 = −ɛ <strong>and</strong> x 3 = 0, determine their subsequentmotions <strong>and</strong> maximum displacements.9.4 Consider the circuit consisting of three equal capacitors <strong>and</strong> two different inductorsshown in the figure. For charges Q i on the capacitors <strong>and</strong> currents I iQ 1 Q 2CCQ 3L 1 C L 2I 1 I 2through the components, write down Kirchhoff’s law <strong>for</strong> the total voltage changearound each of two complete circuit loops. Note that, to within an unimportantconstant, the conservation of current implies that Q 3 = Q 1 − Q 2 . Express the loopequations in the <strong>for</strong>m given in (9.7), namelyA¨Q + BQ = 0.Use this to show that the normal frequencies of the circuit are given byω 2 =1CL 1 L 2[L1 + L 2 ± (L 2 1 + L 2 2 − L 1 L 2 ) 1/2] .Obtain the same matrices <strong>and</strong> result by finding the total energy stored in thevarious capacitors (typically Q 2 /(2C)) <strong>and</strong> in the inductors (typically LI 2 /2).For the special case L 1 = L 2 = L determine the relevant eigenvectors <strong>and</strong> sodescribe the patterns of current flow in the circuit.9.5 It is shown in physics <strong>and</strong> engineering textbooks that circuits containing capacitors<strong>and</strong> inductors can be analysed by replacing a capacitor of capacitance C by a‘complex impedance’ 1/(iωC) <strong>and</strong> an inductor of inductance L by an impedanceiωL, whereω is the angular frequency of the currents flowing <strong>and</strong> i 2 = −1.Use this approach <strong>and</strong> Kirchhoff’s circuit laws to analyse the circuit shown in330


9.4 EXERCISESthe figure <strong>and</strong> obtain three linear equations governing the currents I 1 , I 2 <strong>and</strong> I 3 .Show that the only possible frequencies of self-sustaining currents satisfy eitherPCI 1QLULSI 2 TI 3CCR(a) ω 2 LC =1or(b)3ω 2 LC = 1. Find the corresponding current patterns <strong>and</strong>,in each case, by identifying parts of the circuit in which no current flows, drawan equivalent circuit that contains only one capacitor <strong>and</strong> one inductor.9.6 The simultaneous reduction to diagonal <strong>for</strong>m of two real symmetric quadratic <strong>for</strong>ms.Consider the two real symmetric quadratic <strong>for</strong>ms u T Au <strong>and</strong> u T Bu, whereu Tst<strong>and</strong>s <strong>for</strong> the row matrix (x y z), <strong>and</strong> denote by u n those column matricesthat satisfyBu n = λ n Au n ,(E9.1)in which n is a label <strong>and</strong> the λ n are real, non-zero <strong>and</strong> all different.(a) By multiplying (E9.1) on the left by (u m ) T , <strong>and</strong> the transpose of the correspondingequation <strong>for</strong> u m on the right by u n , show that (u m ) T Au n =0<strong>for</strong>n ≠ m.(b) By noting that Au n =(λ n ) −1 Bu n , deduce that (u m ) T Bu n =0<strong>for</strong>m ≠ n.(c) It can be shown that the u n are linearly independent; the next step is toconstruct a matrix P whose columns are the vectors u n .(d) Make a change of variables u = Pv such that u T Au becomes v T Cv, <strong>and</strong>u T Bubecomes v T Dv. Show that C <strong>and</strong> D are diagonal by showing that c ij =0ifi ≠ j, <strong>and</strong> similarly <strong>for</strong> d ij .Thus u = Pv or v = P −1 u reduces both quadratics to diagonal <strong>for</strong>m.To summarise, the method is as follows:(a) find the λ n that allow (E9.1) a non-zero solution, by solving |B − λA| =0;(b) <strong>for</strong> each λ n construct u n ;(c) construct the non-singular matrix P whose columns are the vectors u n ;(d) make the change of variable u = Pv.9.7 (It is recommended that the reader does not attempt this question until exercise 9.6has been studied.)If, in the pendulum system studied in section 9.1, the string is replaced by asecond rod identical to the first then the expressions <strong>for</strong> the kinetic energy T <strong>and</strong>the potential energy V become(tosecondorderintheθ i )T ≈ Ml ( 2 8 ˙θ 2 3 1 +2˙θ 1˙θ2 + 2 ˙θ 2 3 2),V ≈ Mgl ( 32 θ2 1 + 1 2 2) θ2 .Determine the normal frequencies of the system <strong>and</strong> find new variables ξ <strong>and</strong> ηthat will reduce these two expressions to diagonal <strong>for</strong>m, i.e. toa 1˙ξ2 + a 2˙η 2 <strong>and</strong> b 1 ξ 2 + b 2 η 2 .331


NORMAL MODES9.8 (It is recommended that the reader does not attempt this question until exercise 9.6has been studied.)Find a real linear trans<strong>for</strong>mation that simultaneously reduces the quadratic<strong>for</strong>ms3x 2 +5y 2 +5z 2 +2yz +6zx − 2xy,5x 2 +12y 2 +8yz +4zxto diagonal <strong>for</strong>m.9.9 Three particles of mass m are attached to a light horizontal string having fixedends, the string being thus divided into four equal portions each of length a <strong>and</strong>under a tension T . Show that <strong>for</strong> small transverse vibrations the amplitudes x iof the normal modes satisfy =(maω 2 /T )x, whereB is the matrixBx⎛⎝ −1 2 −12 −1 0⎞⎠ .0 −1 2Estimate the lowest <strong>and</strong> highest eigenfrequencies using trial vectors (3 4 3)(T<strong>and</strong> (3 − 4 3) T . Use also the exact vectors 1 √ ) T (2 1 <strong>and</strong> 1 − √ ) T2 1<strong>and</strong> compare the results.9.10 Use the Rayleigh–Ritz method to estimate the lowest oscillation frequency of aheavy chain of N links, each of length a (= L/N), which hangs freely from oneend. (Try simple calculable configurations such as all links but one vertical, orall links collinear, etc.)9.5 Hints <strong>and</strong> answers9.1 See figure 9.6.9.3 (b) x 1 = ɛ(cos ωt +cos √ 2ωt), x 2 = −ɛ cos √ 2ωt, x 3 = ɛ(− cos ωt +cos √ 2ωt).At various times the three displacements will reach 2ɛ, ɛ, 2ɛ respectively. For example,x 1 canbewrittenas2ɛ cos[( √ 2−1)ωt/2] cos[( √ 2+1)ωt/2], i.e. an oscillationof angular frequency ( √ 2+1)ω/2 <strong>and</strong> modulated amplitude 2ɛ cos[( √ 2−1)ωt/2];the amplitude will reach 2ɛ after a time ≈ 4π/[ω( √ 2 − 1)].9.5 As the circuit loops contain no voltage sources, the equations are homogeneous,<strong>and</strong> so <strong>for</strong> a non-trivial solution the determinant of coefficients must vanish.(a) I 1 =0,I 2 = −I 3 ; no current in PQ; equivalent to two separate circuits ofcapacitance C <strong>and</strong> inductance L.(b) I 1 = −2I 2 = −2I 3 ; no current in TU; capacitance 3C/2 <strong>and</strong> inductance 2L.9.7 ω =(2.634g/l) 1/2 or (0.3661g/l) 1/2 ; θ 1 = ξ + η, θ 2 =1.431ξ − 2.097η.9.9 Estimated, 10/17


9.5 HINTS AND ANSWERS1 2 3mMm(a) ω 2 = c + d mlabel2kM2kmkM(c) ω 2 = c + 2dM + d mFigure 9.6 The normal modes, as viewed from above, of the coupled pendulumsin example 9.1.333


10Vector calculusIn chapter 7 we discussed the algebra of vectors, <strong>and</strong> in chapter 8 we consideredhow to trans<strong>for</strong>m one vector into another using a linear operator. In this chapter<strong>and</strong> the next we discuss the calculus of vectors, i.e. the differentiation <strong>and</strong>integration both of vectors describing particular bodies, such as the velocity ofa particle, <strong>and</strong> of vector fields, in which a vector is defined as a function of thecoordinates throughout some volume (one-, two- or three-dimensional). Since theaim of this chapter is to develop methods <strong>for</strong> h<strong>and</strong>ling multi-dimensional physicalsituations, we will assume throughout that the functions with which we have todeal have sufficiently amenable mathematical properties, in particular that theyare continuous <strong>and</strong> differentiable.10.1 Differentiation of vectorsLet us consider a vector a that is a function of a scalar variable u. By thiswe mean that with each value of u we associate a vector a(u). For example, inCartesian coordinates a(u) =a x (u)i + a y (u)j + a z (u)k, wherea x (u), a y (u) <strong>and</strong>a z (u)are scalar functions of u <strong>and</strong> are the components of the vector a(u) inthex-, y-<strong>and</strong> z- directions respectively. We note that if a(u) is continuous at some pointu = u 0 then this implies that each of the Cartesian components a x (u), a y (u) <strong>and</strong>a z (u) is also continuous there.Let us consider the derivative of the vector function a(u) withrespecttou.The derivative of a vector function is defined in a similar manner to the ordinaryderivative of a scalar function f(x) given in chapter 2. The small change inthe vector a(u) resulting from a small change ∆u in the value of u is given by∆a = a(u +∆u) − a(u) (see figure 10.1). The derivative of a(u) with respect to u isdefined to bedadu = lim∆u→0a(u +∆u) − a(u), (10.1)∆u334


10.1 DIFFERENTIATION OF VECTORSa(u +∆u)∆a = a(u +∆u) − a(u)a(u)Figure 10.1 A small change in a vector a(u) resulting from a small changein u.assuming that the limit exists, in which case a(u) is said to be differentiable atthat point. Note that da/du is also a vector, which is not, in general, parallel toa(u). In Cartesian coordinates, the derivative of the vector a(u) =a x i + a y j + a z kis given bydadu = da xdu i + da ydu j + da zdu k.Perhaps the simplest application of the above is to finding the velocity <strong>and</strong>acceleration of a particle in classical mechanics. If the time-dependent positionvector of the particle with respect to the origin in Cartesian coordinates is givenby r(t) =x(t)i + y(t)j + z(t)k then the velocity of the particle is given by the vectorv(t) = drdt = dxdt i + dydt j + dzdt k.The direction of the velocity vector is along the tangent to the path r(t) attheinstantaneous position of the particle, <strong>and</strong> its magnitude |v(t)| is equal to thespeed of the particle. The acceleration of the particle is given in a similar mannerbya(t) = dvdt = d2 xdt 2 i + d2 ydt 2 j + d2 zdt 2 k.◮The position vector of a particle at time t in Cartesian coordinates is given by r(t) =2t 2 i +(3t − 2)j +(3t 2 − 1)k. Find the speed of the particle at t =1<strong>and</strong> the component ofits acceleration in the direction s = i +2j + k.The velocity <strong>and</strong> acceleration of the particle are given byv(t) = dr =4ti +3j +6tk,dta(t) = dv =4i +6k.dt335


VECTOR CALCULUSyê φjê ρρiφxFigure 10.2coordinates.Unit basis vectors <strong>for</strong> two-dimensional Cartesian <strong>and</strong> plane polarThespeedoftheparticleatt = 1 is simply|v(1)| = √ 4 2 +3 2 +6 2 = √ 61.The acceleration of the particle is constant (i.e. independent of t), <strong>and</strong> its component inthe direction s is given bya · ŝ =(4i +6k) · (i +2j + k)√12 +2 2 +1 2 = 5√ 63 . ◭Note that in the case discussed above i, j <strong>and</strong> k are fixed, time-independentbasis vectors. This may not be true of basis vectors in general; when we arenot using Cartesian coordinates the basis vectors themselves must also be differentiated.We discuss basis vectors <strong>for</strong> non-Cartesian coordinate systems indetail in section 10.10. Nevertheless, as a simple example, let us now considertwo-dimensional plane polar coordinates ρ, φ.Referring to figure 10.2, imagine holding φ fixed <strong>and</strong> moving radially outwards,i.e. in the direction of increasing ρ. Let us denote the unit vector in this directionby ê ρ . Similarly, imagine keeping ρ fixed <strong>and</strong> moving around a circle of fixed radiusin the direction of increasing φ. Let us denote the unit vector tangent to the circleby ê φ . The two vectors ê ρ <strong>and</strong> ê φ are the basis vectors <strong>for</strong> this two-dimensionalcoordinate system, just as i <strong>and</strong> j are basis vectors <strong>for</strong> two-dimensional Cartesiancoordinates. All these basis vectors are shown in figure 10.2.An important difference between the two sets of basis vectors is that, whilei <strong>and</strong> j are constant in magnitude <strong>and</strong> direction, the vectors ê ρ <strong>and</strong> ê φ haveconstant magnitudes but their directions change as ρ <strong>and</strong> φ vary. There<strong>for</strong>e,when calculating the derivative of a vector written in polar coordinates we mustalso differentiate the basis vectors. One way of doing this is to express ê ρ <strong>and</strong> ê φ336


10.1 DIFFERENTIATION OF VECTORSin terms of i <strong>and</strong> j. From figure 10.2, we see thatê ρ =cosφ i +sinφ j,ê φ = − sin φ i +cosφ j.Since i <strong>and</strong> j are constant vectors, we find that the derivatives of the basis vectorsê ρ <strong>and</strong> ê φ with respect to t are given bydê ρdtdê φdt= − sin φ dφdt i +cosφdφ dt j = ˙φ ê φ , (10.2)= − cos φ dφ i − sin φdφdt dt j = −˙φ ê ρ , (10.3)where the overdot is the conventional notation <strong>for</strong> differentiation with respect totime.◮The position vector of a particle in plane polar coordinates is r(t) =ρ(t)ê ρ .Findexpressions<strong>for</strong> the velocity <strong>and</strong> acceleration of the particle in these coordinates.Using result (10.4) below, the velocity of the particle is given byv(t) =ṙ(t) =˙ρ ê ρ + ρ ˙ê ρ = ˙ρ ê ρ + ρ˙φ ê φ ,where we have used (10.2). In a similar way its acceleration is given bya(t) = d dt (˙ρ ê ρ + ρ˙φ ê φ )= ¨ρ ê ρ + ˙ρ ˙ê ρ + ρ˙φ ˙ê φ + ρ¨φ ê φ + ˙ρ˙φ ê φ= ¨ρ ê ρ + ˙ρ(˙φ ê φ )+ρ˙φ(−˙φ ê ρ )+ρ¨φ ê φ + ˙ρ˙φ ê φ=(¨ρ − ρ˙φ 2 ) ê ρ +(ρ¨φ +2˙ρ˙φ) ê φ . ◭Here we have used (10.2) <strong>and</strong> (10.3).10.1.1 Differentiation of composite vector expressionsIn composite vector expressions each of the vectors or scalars involved may bea function of some scalar variable u, as we have seen. The derivatives of suchexpressions are easily found using the definition (10.1) <strong>and</strong> the rules of ordinarydifferential calculus. They may be summarised by the following, in which weassume that a <strong>and</strong> b are differentiable vector functions of a scalar u <strong>and</strong> that φis a differentiable scalar function of u:d(φa) =φdadu du + dφ a, (10.4)duddb(a · b) =a ·du du + da · b, (10.5)duddb(a × b) =a ×du du + da × b. (10.6)du337


VECTOR CALCULUSThe order of the factors in the terms on the RHS of (10.6) is, of course, just asimportant as it is in the original vector product.◮A particleofmassm with position vector r relative to some origin O experiences a <strong>for</strong>ceF, which produces a torque (moment) T = r × F about O. The angular momentum of theparticle about O is given by L = r × mv, wherev is the particle’s velocity. Show that therate of change of angular momentum is equal to the applied torque.The rate of change of angular momentum is given bydLdt = d (r × mv).dtUsing (10.6) we obtaindLdt = drdt × mv + r × d dt (mv)= v × mv + r × d dt (mv)= 0 + r × F = T,where in the last line we use Newton’s second law, namely F = d(mv)/dt. ◭If a vector a is a function of a scalar variable s that is itself a function of u, sothat s = s(u), then the chain rule (see subsection 2.1.3) givesda(s)du= ds dadu ds . (10.7)The derivatives of more complicated vector expressions may be found by repeatedapplication of the above equations.One further useful result can be derived by considering the derivativedda(a · a) =2a ·du du ;since a · a = a 2 ,wherea = |a|, weseethata · da =0 ifa is constant. (10.8)duIn other words, if a vector a(u) has a constant magnitude as u varies then it isperpendicular to the vector da/du.10.1.2 Differential of a vectorAs a final note on the differentiation of vectors, we can also define the differentialof a vector, in a similar way to that of a scalar in ordinary differential calculus.In the definition of the vector derivative (10.1), we used the notion of a smallchange ∆a in a vector a(u) resulting from a small change ∆u in its argument. Inthe limit ∆u → 0, the change in a becomes infinitesimally small, <strong>and</strong> we denote itby the differential da. From (10.1) we see that the differential is given byda = da du. (10.9)du338


10.2 INTEGRATION OF VECTORSNote that the differential of a vector is also a vector. As an example, theinfinitesimal change in the position vector of a particle in an infinitesimal time dt isdr = dr dt = v dt,dtwhere v is the particle’s velocity.10.2 Integration of vectorsThe integration of a vector (or of an expression involving vectors that may itselfbe either a vector or scalar) with respect to a scalar u canberegardedastheinverse of differentiation. We must remember, however, that(i) the integral has the same nature (vector or scalar) as the integr<strong>and</strong>,(ii) the constant of integration <strong>for</strong> indefinite integrals must be of the samenature as the integral.For example, if a(u) =d[A(u)]/du then the indefinite integral of a(u) is given by∫a(u) du = A(u)+b,where b is a constant vector. The definite integral of a(u) fromu = u 1 to u = u 2is given by∫ u2u 1a(u) du = A(u 2 ) − A(u 1 ).◮A small particle of mass m orbits a much larger mass M centred at the origin O. Accordingto Newton’s law of gravitation, the position vector r of the small mass obeys the differentialequationm d2 rdt = − GMm ˆr.2 r 2Show that the vector r × dr/dt is a constant of the motion.Forming the vector product of the differential equation with r, weobtainr × d2 rdt = − GM2 r r × ˆr.2Since r <strong>and</strong> ˆr are collinear, r × ˆr = 0 <strong>and</strong> there<strong>for</strong>e we haveHowever,r × d2 r= 0. (10.10)dt2 (dr × dr )= r × d2 rdt dt dt + dr2 dt × drdt = 0,339


VECTOR CALCULUSzCˆnPˆtˆbr(u)OyxFigure 10.3 The unit tangent ˆt, normal ˆn <strong>and</strong> binormal ˆb to the space curve Cat a particular point P .since the first term is zero by (10.10), <strong>and</strong> the second is zero because it is the vector productof two parallel (in this case identical) vectors. Integrating, we obtain the required resultr × dr = c, (10.11)dtwhere c is a constant vector.As a further point of interest we may note that in an infinitesimal time dt the changein the position vector of the small mass is dr <strong>and</strong> the element of area swept out by theposition vector of the particle is simply dA = 1 |r×dr|. Dividing both sides of this equation2by dt, we conclude thatdAdt = 1 2 ∣ r × drdt ∣ = |c|2 ,<strong>and</strong> that the physical interpretation of the above result (10.11) is that the position vector rof the small mass sweeps out equal areas in equal times. This result is in fact valid <strong>for</strong>motion under any <strong>for</strong>ce that acts along the line joining the two particles. ◭10.3 Space curvesIn the previous section we mentioned that the velocity vector of a particle is atangent to the curve in space along which the particle moves. We now give a morecomplete discussion of curves in space <strong>and</strong> also a discussion of the geometricalinterpretation of the vector derivative.AcurveC in space can be described by the vector r(u) joining the origin O ofa coordinate system to a point on the curve (see figure 10.3). As the parameter uvaries, the end-point of the vector moves along the curve. In Cartesian coordinates,r(u) =x(u)i + y(u)j + z(u)k,where x = x(u), y = y(u) <strong>and</strong>z = z(u) aretheparametric equations of the curve.340


10.3 SPACE CURVESThis parametric representation can be very useful, particularly in mechanics whenthe parameter may be the time t. We can, however, also represent a space curveby y = f(x), z = g(x), which can be easily converted into the above parametric<strong>for</strong>m by setting u = x, sothatr(u) =ui + f(u)j + g(u)k.Alternatively, a space curve can be represented in the <strong>for</strong>m F(x, y, z) = 0,G(x, y, z) = 0, where each equation represents a surface <strong>and</strong> the curve is theintersection of the two surfaces.A curve may sometimes be described in parametric <strong>for</strong>m by the vector r(s),where the parameter s is the arc length along the curve measured from a fixedpoint. Even when the curve is expressed in terms of some other parameter, it isstraight<strong>for</strong>ward to find the arc length between any two points on the curve. Forthe curve described by r(u), let us consider an infinitesimal vector displacementdr = dx i + dy j + dz kalong the curve. The square of the infinitesimal distance moved is then given by(ds) 2 = dr · dr =(dx) 2 +(dy) 2 +(dz) 2 ,fromwhichitcanbeshownthat( ) ds 2= drdu du · drdu .There<strong>for</strong>e, the arc length between two points on the curve r(u), given by u = u 1<strong>and</strong> u = u 2 ,iss =∫ u2u 1√drdu · dr du. (10.12)du◮ Acurvelyinginthexy-plane is given by y = y(x), z =0. Using (10.12), show that thearc length along the curve between x = a <strong>and</strong> x = b is given by s = ∫ √ba 1+y ′2 dx, wherey ′ = dy/dx.Let us first represent the curve in parametric <strong>for</strong>m by setting u = x, sothatr(u) =ui + y(u)j.Differentiating with respect to u, we finddrdu = i + dydu j,from which we obtaindrdu · dr ( ) 2 dydu =1+ .du341


VECTOR CALCULUSThere<strong>for</strong>e, remembering that u = x, from (10.12) the arc length between x = a <strong>and</strong> x = bis given by∫ b√ √drs =a du · dr ∫ b( ) 2 dydu du = 1+ dx.a dxThis result was derived using more elementary methods in chapter 2. ◭If a curve C is described by r(u) then, by considering figures 10.1 <strong>and</strong> 10.3, wesee that, at any given point on the curve, dr/du is a vector tangent to C at thatpoint, in the direction of increasing u. In the special case where the parameter uis the arc length s along the curve then dr/ds is a unit tangent vector to C <strong>and</strong> isdenoted by ˆt.The rate at which the unit tangent ˆt changes with respect to s is given byd ˆt/ds, <strong>and</strong> its magnitude is defined as the curvature κ of the curve C at a givenpoint,∣ ∣ κ =d ˆt∣∣∣ ∣ ds ∣ = d 2 ˆr ∣∣∣ds 2 .We can also define the quantity ρ =1/κ, which is called the radius of curvature.Since ˆt is of constant (unit) magnitude, it follows from (10.8) that it is perpendicularto d ˆt/ds. The unit vector in the direction perpendicular to ˆt is denotedby ˆn <strong>and</strong> is called the principal normal at the point. We there<strong>for</strong>e haved ˆt= κ ˆn. (10.13)dsThe unit vector ˆb = ˆt × ˆn, which is perpendicular to the plane containing ˆt<strong>and</strong> ˆn, is called the binormal to C. The vectors ˆt, ˆn <strong>and</strong> ˆb <strong>for</strong>m a right-h<strong>and</strong>edrectangular cooordinate system (or triad) at any given point on C (see figure 10.3).As s changes so that the point of interest moves along C, the triad of vectors alsochanges.The rate at which ˆb changes with respect to s is given by d ˆb/ds <strong>and</strong> is ameasure of the torsion τ of the curve at any given point. Since ˆb is of constantmagnitude, from (10.8) it is perpendicular to d ˆb/ds. We may further show thatd ˆb/ds is also perpendicular to ˆt, as follows. By definition ˆb · ˆt =0,whichondifferentiating yields0= d ( )ˆb · ˆt = d ˆbds ds · ˆt + ˆb · d ˆtds= d ˆbds · ˆt + ˆb · κ ˆn= d ˆbds · ˆt,where we have used the fact that ˆb · ˆn = 0. Hence, since d ˆb/ds is perpendicularto both ˆb <strong>and</strong> ˆt, we must have d ˆb/ds ∝ ˆn. The constant of proportionality is −τ,342


10.3 SPACE CURVESso we finally obtaind ˆb = −τ ˆn. (10.14)dsTaking the dot product of each side with ˆn, we see that the torsion of a curve isgiven byτ = − ˆn · d ˆbds .We may also define the quantity σ =1/τ, which is called the radius of torsion.Finally, we consider the derivative d ˆn/ds. Since ˆn = ˆb × ˆt we haved ˆnds = d ˆbds × ˆt + ˆb × d ˆtds= −τ ˆn × ˆt + ˆb × κ ˆn= τ ˆb − κ ˆt. (10.15)In summary, ˆt, ˆn <strong>and</strong> ˆb <strong>and</strong> their derivatives with respect to s are related to oneanother by the relations (10.13), (10.14) <strong>and</strong> (10.15), the Frenet–Serret <strong>for</strong>mulae,d ˆtds = κ ˆn,d ˆnds = τ ˆb − κ ˆt,d ˆbds= −τ ˆn. (10.16)◮Show that the acceleration of a particle travelling along a trajectory r(t) is given bya(t) = dvdt ˆt + v2ρ ˆn,where v is the speed of the particle, ˆt is the unit tangent to the trajectory, ˆn is its principalnormal <strong>and</strong> ρ is its radius of curvature.The velocity of the particle is given byv(t) = drdt = dr dsds dt = dsdt ˆt,where ds/dt is the speed of the particle, which we denote by v, <strong>and</strong> ˆt is the unit vectortangent to the trajectory. Writing the velocity as v = v ˆt, <strong>and</strong> differentiating once morewith respect to time t, weobtaina(t) = dvdt = dvdt ˆt + v d ˆtdt ;but we note thatd ˆtdt = ds d ˆtdt ds = vκ ˆn = v ρ ˆn.There<strong>for</strong>e, we havea(t) = dvdt ˆt + v2ρ ˆn.This shows that in addition to an acceleration dv/dt along the tangent to the particle’strajectory, there is also an acceleration v 2 /ρ in the direction of the principal normal. Thelatter is often called the centripetal acceleration. ◭343


VECTOR CALCULUSFinally, we note that a curve r(u) representing the trajectory of a particle maysometimes be given in terms of some parameter u that is not necessarily equal tothe time t but is functionally related to it in some way. In this case the velocityof the particle is given byv = drdt = dr dudu dt .Differentiating again with respect to time gives the acceleration asa = dvdt = d ( ) ( ) 2 dr du= d2 r dudt du dt du 2 + dr d 2 udt du dt 2 .10.4 Vector functions of several argumentsThe concept of the derivative of a vector is easily extended to cases where thevectors (or scalars) are functions of more than one independent scalar variable,u 1 ,u 2 ,...,u n . In this case, the results of subsection 10.1.1 are still valid, exceptthat the derivatives become partial derivatives ∂a/∂u i defined as in ordinarydifferential calculus. For example, in Cartesian coordinates,∂a∂u = ∂a x∂u i + ∂a y∂u j + ∂a z∂u k.In particular, (10.7) generalises to the chain rule of partial differentiation discussedin section 5.5. If a = a(u 1 ,u 2 ,...,u n ) <strong>and</strong> each of the u i is also a functionu i (v 1 ,v 2 ,...,v n ) of the variables v i then, generalising (5.17),∂a∂v i= ∂a∂u 1∂u 1∂v i+ ∂a∂u 2∂u 2∂v i+ ···+ ∂a∂u n∂u n∂v i=n∑j=1∂a∂u j∂u j∂v i. (10.17)A special case of this rule arises when a is an explicit function of some variablev, as well as of scalars u 1 ,u 2 ,...,u n that are themselves functions of v; thenwehavedadv = ∂a∂v +n∑j=1∂a∂u j∂u j∂v . (10.18)We may also extend the concept of the differential of a vector given in (10.9)to vectors dependent on several variables u 1 ,u 2 ,...,u n :da = ∂a du 1 + ∂a du 2 + ···+ ∂an∑ ∂adu n = du j . (10.19)∂u 1 ∂u 2 ∂u n ∂u jAs an example, the infinitesimal change in an electric field E in moving from aposition r to a neighbouring one r + dr is given bydE = ∂E∂xdx +∂E∂y344j=1∂Edy + dz. (10.20)∂z


10.5 SURFACESzT∂r∂uSu = c 1P∂r∂vr(u, v)v = c 2OyxFigure 10.4 The tangent plane T to a surface S at a particular point P ;u = c 1 <strong>and</strong> v = c 2 are the coordinate curves, shown by dotted lines, that passthrough P . The broken line shows some particular parametric curve r = r(λ)lying in the surface.10.5 SurfacesAsurfaceS in space can be described by the vector r(u, v) joining the origin O ofa coordinate system to a point on the surface (see figure 10.4). As the parametersu <strong>and</strong> v vary, the end-point of the vector moves over the surface. This is verysimilar to the parametric representation r(u) of a curve, discussed in section 10.3,but with the important difference that we require two parameters to describe asurface, whereas we need only one to describe a curve.In Cartesian coordinates the surface is given byr(u, v) =x(u, v)i + y(u, v)j + z(u, v)k,where x = x(u, v), y = y(u, v) <strong>and</strong>z = z(u, v) are the parametric equations of thesurface. We can also represent a surface by z = f(x, y) org(x, y, z) =0.Eitherof these representations can be converted into the parametric <strong>for</strong>m in a similarmanner to that used <strong>for</strong> equations of curves. For example, if z = f(x, y) thenbysetting u = x <strong>and</strong> v = y the surface can be represented in parametric <strong>for</strong>m byr(u, v) =ui + vj + f(u, v)k.Any curve r(λ), where λ is a parameter, on the surface S can be representedby a pair of equations relating the parameters u <strong>and</strong> v, <strong>for</strong> example u = f(λ)<strong>and</strong> v = g(λ). A parametric representation of the curve can easily be found bystraight<strong>for</strong>ward substitution, i.e. r(λ) =r(u(λ),v(λ)). Using (10.17) <strong>for</strong> the casewhere the vector is a function of a single variable λ so that the LHS becomes a345


VECTOR CALCULUStotal derivative, the tangent to the curve r(λ) at any point is given bydrdλ = ∂r du∂u dλ + ∂r dv∂v dλ . (10.21)The two curves u =constant<strong>and</strong>v = constant passing through any point Pon S are called coordinate curves. Forthecurveu = constant, <strong>for</strong> example, wehave du/dλ = 0, <strong>and</strong> so from (10.21) its tangent vector is in the direction ∂r/∂v.Similarly, the tangent vector to the curve v = constant is in the direction ∂r/∂u.If the surface is smooth then at any point P on S the vectors ∂r/∂u <strong>and</strong>∂r/∂v are linearly independent <strong>and</strong> define the tangent plane T at the point P (seefigure 10.4). A vector normal to the surface at P is given byn = ∂r∂u × ∂r∂v . (10.22)In the neighbourhood of P , an infinitesimal vector displacement dr is writtendr = ∂r ∂rdu +∂u ∂v dv.The element of area at P , an infinitesimal parallelogram whose sides are thecoordinate curves, has magnitude∣ dS =∂r ∂r ∣∣∣∣ du ×∂u ∂v dv =∂r∣∂u × ∂r∂v ∣ du dv = |n| du dv. (10.23)Thus the total area of the surface is∫∫A =∂r∣∂u ∫∫∂v∣ |n| du dv, (10.24)Rwhere R is the region in the uv-plane corresponding to the range of parametervalues that define the surface.◮ Find the element of area on the surface of a sphere of radius a, <strong>and</strong> hence calculate thetotal surface area of the sphere.We can represent a point r on the surface of the sphere in terms of the two parameters θ<strong>and</strong> φ:r(θ, φ) =a sin θ cos φ i + a sin θ sin φ j + a cos θ k,where θ <strong>and</strong> φ are the polar <strong>and</strong> azimuthal angles respectively. At any point P ,vectorstangent to the coordinate curves θ =constant<strong>and</strong>φ = constant are∂r= a cos θ cos φ i + a cos θ sin φ j − a sin θ k,∂θ∂r= −a sin θ sin φ i + a sin θ cos φ j.∂φ346R


10.6 SCALAR AND VECTOR FIELDSA normal n to the surface at this point is then given by∣ ∣∣∣∣∣∣n = ∂r∂θ × ∂ri j k∂φ = a cos θ cos φ acos θ sin φ −a sin θ−a sin θ sin φ asin θ cos φ 0∣= a 2 sin θ(sin θ cos φ i +sinθ sin φ j +cosθ k),which has a magnitude of a 2 sin θ. There<strong>for</strong>e, the element of area at P is, from (10.23),dS = a 2 sin θdθdφ,<strong>and</strong> the total surface area of the sphere is given byA =∫ πdθ∫ 2π0 0dφ a 2 sin θ =4πa 2 .This familiar result can, of course, be proved by much simpler methods! ◭10.6 Scalar <strong>and</strong> vector fieldsWe now turn to the case where a particular scalar or vector quantity is definednot just at a point in space but continuously as a field throughout some regionof space R (which is often the whole space). Although the concept of a field isvalid <strong>for</strong> spaces with an arbitrary number of dimensions, in the remainder of thischapter we will restrict our attention to the familiar three-dimensional case. Ascalar field φ(x, y, z) associates a scalar with each point in R, while a vector fielda(x, y, z) associates a vector with each point. In what follows, we will assume thatthe variation in the scalar or vector field from point to point is both continuous<strong>and</strong> differentiable in R.Simple examples of scalar fields include the pressure at each point in a fluid<strong>and</strong> the electrostatic potential at each point in space in the presence of an electriccharge. Vector fields relating to the same physical systems are the velocity vectorin a fluid (giving the local speed <strong>and</strong> direction of the flow) <strong>and</strong> the electric field.With the study of continuously varying scalar <strong>and</strong> vector fields there arises theneed to consider their derivatives <strong>and</strong> also the integration of field quantities alonglines, over surfaces <strong>and</strong> throughout volumes in the field. We defer the discussionof line, surface <strong>and</strong> volume integrals until the next chapter, <strong>and</strong> in the remainderof this chapter we concentrate on the definition of vector differential operators<strong>and</strong> their properties.10.7 Vector operatorsCertain differential operations may be per<strong>for</strong>med on scalar <strong>and</strong> vector fields<strong>and</strong> have wide-ranging applications in the physical sciences. The most importantoperations are those of finding the gradient of a scalar field <strong>and</strong> the divergence<strong>and</strong> curl of a vector field. It is usual to define these operators from a strictly347


VECTOR CALCULUSmathematical point of view, as we do below. In the following chapter, however, wewill discuss their geometrical definitions, which rely on the concept of integratingvector quantities along lines <strong>and</strong> over surfaces.Central to all these differential operations is the vector operator ∇, whichiscalled del (or sometimes nabla) <strong>and</strong> in Cartesian coordinates is defined by∇≡i ∂∂x + j ∂∂y + k ∂ ∂z . (10.25)The <strong>for</strong>m of this operator in non-Cartesian coordinate systems is discussed insections 10.9 <strong>and</strong> 10.10.10.7.1 Gradient of a scalar fieldThe gradient of a scalar field φ(x, y, z) is defined bygrad φ = ∇φ = i ∂φ∂x + j∂φ ∂y + k∂φ ∂z . (10.26)Clearly, ∇φ is a vector field whose x-, y- <strong>and</strong>z- components are the first partialderivatives of φ(x, y, z) with respect to x, y <strong>and</strong> z respectively. Also note that thevector field ∇φ should not be confused with the vector operator φ∇, which hascomponents (φ ∂/∂x, φ ∂/∂y, φ ∂/∂z).◮Find the gradient of the scalar field φ = xy 2 z 3 .From (10.26) the gradient of φ is given by∇φ = y 2 z 3 i +2xyz 3 j +3xy 2 z 2 k. ◭The gradient of a scalar field φ has some interesting geometrical properties.Let us first consider the problem of calculating the rate of change of φ in someparticular direction. For an infinitesimal vector displacement dr, <strong>for</strong>ming its scalarproduct with ∇φ we obtain(∇φ · dr = i ∂φ)∂x + j∂φ ∂y + k∂φ · (i dx + j dy + k dx) ,∂z= ∂φ ∂φ ∂φdx + dy +∂x ∂y ∂z dz,= dφ, (10.27)which is the infinitesimal change in φ in going from position r to r + dr. Inparticular, if r depends on some parameter u such that r(u) defines a space curve348


10.7 VECTOR OPERATORS∇φaθQPdφdsin the direction aφ =constantFigure 10.5 Geometrical properties of ∇φ. PQ gives the value of dφ/ds inthe direction a.then the total derivative of φ with respect to u along the curve is simplydφ dr= ∇φ ·du du . (10.28)In the particular case where the parameter u is the arc length s along the curve,the total derivative of φ with respect to s along the curve is given bydφds = ∇φ · ˆt, (10.29)where ˆt is the unit tangent to the curve at the given point, as discussed insection 10.3.In general, the rate of change of φ with respect to the distance s in a particulardirection a is given bydφ= ∇φ · â (10.30)ds<strong>and</strong> is called the directional derivative. Since â is a unit vector we havedφ= |∇φ| cos θdswhere θ is the angle between â <strong>and</strong> ∇φ as shown in figure 10.5. Clearly ∇φ liesin the direction of the fastest increase in φ, <strong>and</strong>|∇φ| is the largest possible valueof dφ/ds. Similarly, the largest rate of decrease of φ is dφ/ds = −|∇φ| in thedirection of −∇φ.349


VECTOR CALCULUS◮For the function φ = x 2 y + yz at the point (1, 2, −1), find its rate of change with distancein the direction a = i +2j +3k. At this same point, what is the greatest possible rate ofchange with distance <strong>and</strong> in which direction does it occur?The gradient of φ is given by (10.26):∇φ =2xyi +(x 2 + z)j + yk,=4i +2k at the point (1, 2, −1).The unit vector in the direction of a is â = √ 114(i +2j +3k), so the rate of change of φwith distance s in this direction is, using (10.30),dφds = ∇φ · â = √ 1 (4+6)= √ 10 .14 14From the above discussion, at the point (1, 2, −1) dφ/ds will be greatest in the directionof ∇φ =4i +2k <strong>and</strong> has the value |∇φ| = √ 20 in this direction. ◭We can extend the above analysis to find the rate of change of a vectorfield (rather than a scalar field as above) in a particular direction. The scalardifferential operator â · ∇ can be shown to give the rate of change with distancein the direction â of the quantity (vector or scalar) on which it acts. In Cartesiancoordinates it may be written as∂â · ∇ = a x∂x + a ∂y∂y + a ∂z∂z . (10.31)Thus we can write the infinitesimal change in an electric field in moving from rto r + dr given in (10.20) as dE =(dr · ∇)E.A second interesting geometrical property of ∇φ may be found by consideringthe surface defined by φ(x, y, z) =c, wherec is some constant. If ˆt is a unittangent to this surface at some point then clearly dφ/ds = 0 in this direction<strong>and</strong> from (10.29) we have ∇φ · ˆt = 0. In other words, ∇φ is a vector normal tothe surface φ(x, y, z) =c at every point, as shown in figure 10.5. If ˆn is a unitnormal to the surface in the direction of increasing φ(x, y, z), then the gradient issometimes written∇φ ≡ ∂φ ˆn, (10.32)∂nwhere ∂φ/∂n ≡|∇φ| is the rate of change of φ in the direction ˆn <strong>and</strong> is calledthe normal derivative.◮Find expressions <strong>for</strong> the equations of the tangent plane <strong>and</strong> the line normal to the surfaceφ(x, y, z) =c at the point P with coordinates x 0 ,y 0 ,z 0 . Use the results to find the equationsof the tangent plane <strong>and</strong> the line normal to the surface of the sphere φ = x 2 + y 2 + z 2 = a 2at the point (0, 0,a).A vector normal to the surface φ(x, y, z) =c at the point P is simply ∇φ evaluated at thatpoint; we denote it by n 0 .Ifr 0 is the position vector of the point P relative to the origin,350


10.7 VECTOR OPERATORSz(0, 0,a)ˆn 0z = aOayφ = x 2 + y 2 + z 2 = a 2xFigure 10.6 The tangent plane <strong>and</strong> the normal to the surface of the sphereφ = x 2 + y 2 + z 2 = a 2 at the point r 0 with coordinates (0, 0,a).<strong>and</strong> r is the position vector of any point on the tangent plane, then the vector equation ofthe tangent plane is, from (7.41),(r − r 0 ) · n 0 =0.Similarly, if r is the position vector of any point on the straight line passing through P(with position vector r 0 ) in the direction of the normal n 0 then the vector equation of thisline is, from subsection 7.7.1,(r − r 0 ) × n 0 = 0.For the surface of the sphere φ = x 2 + y 2 + z 2 = a 2 ,∇φ =2xi +2yj +2zk=2ak at the point (0, 0,a).There<strong>for</strong>e the equation of the tangent plane to the sphere at this point is(r − r 0 ) · 2ak =0.This gives 2a(z − a) =0orz = a, as expected. The equation of the line normal to thesphere at the point (0, 0,a)is(r − r 0 ) × 2ak = 0,which gives 2ayi − 2axj = 0 or x = y =0,i.e.thez-axis, as expected. The tangent plane<strong>and</strong> normal to the surface of the sphere at this point are shown in figure 10.6. ◭Further properties of the gradient operation, which are analogous to those ofthe ordinary derivative, are listed in subsection 10.8.1 <strong>and</strong> may be easily proved.351


VECTOR CALCULUSIn addition to these, we note that the gradient operation also obeys the chainrule as in ordinary differential calculus, i.e. if φ <strong>and</strong> ψ are scalar fields in someregion R then∇ [φ(ψ)] = ∂φ∂ψ ∇ψ.10.7.2 Divergence of a vector fieldThe divergence of a vector field a(x, y, z) is defined bydiv a = ∇ · a = ∂a x∂x + ∂a y∂y + ∂a z∂z , (10.33)where a x , a y <strong>and</strong> a z are the x-, y- <strong>and</strong>z- components of a. Clearly, ∇ · a is a scalarfield. Any vector field a <strong>for</strong> which ∇ · a = 0 is said to be solenoidal.◮Find the divergence of the vector field a = x 2 y 2 i + y 2 z 2 j + x 2 z 2 k.From (10.33) the divergence of a is given by∇ · a =2xy 2 +2yz 2 +2x 2 z =2(xy 2 + yz 2 + x 2 z). ◭We will discuss fully the geometric definition of divergence <strong>and</strong> its physicalmeaning in the next chapter. For the moment, we merely note that the divergencecan be considered as a quantitative measure of how much a vector field diverges(spreads out) or converges at any given point. For example, if we consider thevector field v(x, y, z) describing the local velocity at any point in a fluid then ∇ · vis equal to the net rate of outflow of fluid per unit volume, evaluated at a point(by letting a small volume at that point tend to zero).Now if some vector field a is itself derived from a scalar field via a = ∇φ then∇ · a has the <strong>for</strong>m ∇ · ∇φ or, as it is usually written, ∇ 2 φ,where∇ 2 (del squared)is the scalar differential operator∇ 2 ≡ ∂2∂x 2 + ∂2∂y 2 + ∂2∂z 2 . (10.34)∇ 2 φ is called the Laplacian of φ <strong>and</strong> appears in several important partial differentialequations of mathematical physics, discussed in chapters 20 <strong>and</strong> 21.◮Find the Laplacian of the scalar field φ = xy 2 z 3 .From (10.34) the Laplacian of φ is given by∇ 2 φ = ∂2 φ∂x + ∂2 φ2 ∂y + ∂2 φ2 ∂z 2 =2xz3 +6xy 2 z. ◭352


10.7 VECTOR OPERATORS10.7.3 Curl of a vector fieldThe curl of a vector field a(x, y, z) is defined by( ∂azcurl a = ∇×a =∂y − ∂a ) (y ∂axi +∂z ∂z − ∂a ) (z ∂ayj +∂x ∂x − ∂a )xk,∂ywhere a x , a y <strong>and</strong> a z are the x-, y- <strong>and</strong>z- components of a. The RHS can bewritten in a more memorable <strong>for</strong>m as a determinant:∣ i j k ∣∣∣∣∣∣∣ ∇×a =∂ ∂ ∂, (10.35)∂x ∂y ∂z∣ a x a y a zwhere it is understood that, on exp<strong>and</strong>ing the determinant, the partial derivativesin the second row act on the components of a in the third row. Clearly, ∇×a isitself a vector field. Any vector field a <strong>for</strong> which ∇×a = 0 is said to be irrotational.◮Find the curl of the vector field a = x 2 y 2 z 2 i + y 2 z 2 j + x 2 z 2 k.The curl of a is given by∣ i j k ∣∣∣∣∣∣∣ ∂ ∂ ∂∇φ == −2 [ y 2 zi +(xz 2 − x 2 y 2 z)j + x 2 yz 2 k ] . ◭∂x ∂y ∂z∣x 2 y 2 z 2 y 2 z 2 x 2 z 2For a vector field v(x, y, z) describing the local velocity at any point in a fluid,∇×v is a measure of the angular velocity of the fluid in the neighbourhood ofthat point. If a small paddle wheel were placed at various points in the fluid thenit would tend to rotate in regions where ∇×v ≠ 0, while it would not rotate inregions where ∇×v = 0.Another insight into the physical interpretation of the curl operator is gainedby considering the vector field v describing the velocity at any point in a rigidbody rotating about some axis with angular velocity ω. Ifr is the position vectorof the point with respect to some origin on the axis of rotation then the velocityof the point is given by v = ω × r. Without any loss of generality, we may takeω to lie along the z-axis of our coordinate system, so that ω = ω k. The velocityfield is then v = −ωy i + ωx j. The curl of this vector field is easily found to bei j k∇×v =∂ ∂ ∂∂x ∂y ∂z=2ωk =2ω. (10.36)∣ −ωy ωx 0 ∣353


VECTOR CALCULUS∇(φ + ψ) =∇φ + ∇ψ∇ · (a + b) =∇ · a + ∇ · b∇×(a + b) =∇×a + ∇×b∇(φψ) =φ∇ψ + ψ∇φ∇(a · b) =a × (∇×b)+b × (∇×a)+(a · ∇)b +(b · ∇)a∇ · (φa) =φ∇ · a + a · ∇φ∇ · (a × b) =b · (∇×a) − a · (∇×b)∇×(φa) =∇φ × a + φ∇×a∇×(a × b) =a(∇ · b) − b(∇ · a)+(b · ∇)a − (a · ∇)bTable 10.1 Vector operators acting on sums <strong>and</strong> products. The operator ∇ isdefined in (10.25); φ <strong>and</strong> ψ are scalar fields, a <strong>and</strong> b are vector fields.There<strong>for</strong>e the curl of the velocity field is a vector equal to twice the angularvelocity vector of the rigid body about its axis of rotation. We give a fullgeometrical discussion of the curl of a vector in the next chapter.10.8 Vector operator <strong>for</strong>mulaeIn the same way as <strong>for</strong> ordinary vectors (chapter 7), <strong>for</strong> vector operators certainidentities exist. In addition, we must consider various relations involving theaction of vector operators on sums <strong>and</strong> products of scalar <strong>and</strong> vector fields. Someof these relations have been mentioned earlier, but we list all the most importantones here <strong>for</strong> convenience. The validity of these relations may be easily verifiedby direct calculation (a quick method of deriving them using tensor notation isgiven in chapter 26).Although some of the following vector relations are expressed in Cartesiancoordinates, it may be proved that they are all independent of the choice ofcoordinate system. This is to be expected since grad, div <strong>and</strong> curl all have cleargeometrical definitions, which are discussed more fully in the next chapter <strong>and</strong>which do not rely on any particular choice of coordinate system.10.8.1 Vector operators acting on sums <strong>and</strong> productsLet φ <strong>and</strong> ψ be scalar fields <strong>and</strong> a <strong>and</strong> b be vector fields. Assuming these fieldsare differentiable, the action of grad, div <strong>and</strong> curl on various sums <strong>and</strong> productsof them is presented in table 10.1.These relations can be proved by direct calculation.354


10.8 VECTOR OPERATOR FORMULAE◮Show that∇×(φa) =∇φ × a + φ∇×a.The x-component of the LHS is∂∂y (φa z) − ∂ ∂z (φa y)=φ ∂a z∂y + ∂φ∂y a z − φ ∂a y∂z − ∂φ∂z a y,( ∂az= φ∂y − ∂a ) (y ∂φ+∂z ∂y a z − ∂φ )∂z a y ,= φ(∇×a) x +(∇φ × a) x ,where, <strong>for</strong> example, (∇φ×a) x denotes the x-component of the vector ∇φ×a. Incorporatingthe y- <strong>and</strong>z- components, which can be similarly found, we obtain the stated result. ◭Some useful special cases of the relations in table 10.1 are worth noting. If r isthe position vector relative to some origin <strong>and</strong> r = |r|, then∇φ(r) = dφdr ˆr,∇ · [φ(r)r] =3φ(r)+r dφ(r) ,dr∇ 2 φ(r) = d2 φ(r)dr 2 + 2 dφ(r),r dr∇×[φ(r)r] = 0.These results may be proved straight<strong>for</strong>wardly using Cartesian coordinates butfar more simply using spherical polar coordinates, which are discussed in subsection10.9.2. Particular cases of these results aretogether with∇r = ˆr, ∇ · r =3, ∇×r = 0,( ) 1∇ = − ˆrr r 2 ,( ) ( ˆr1∇ ·r 2 = −∇ 2 r)=4πδ(r),where δ(r) is the Dirac delta function, discussed in chapter 13. The last equation isimportant in the solution of certain partial differential equations <strong>and</strong> is discussedfurther in chapter 20.10.8.2 Combinations of grad, div <strong>and</strong> curlWe now consider the action of two vector operators in succession on a scalar orvector field. We can immediately discard four of the nine obvious combinations ofgrad, div <strong>and</strong> curl, since they clearly do not make sense. If φ is a scalar field <strong>and</strong>355


VECTOR CALCULUSa is a vector field, these four combinations are grad(grad φ), div(div a), curl(div a)<strong>and</strong> grad(curl a). In each case the second (outer) vector operator is acting on thewrong type of field, i.e. scalar instead of vector or vice versa. In grad(grad φ), <strong>for</strong>example, grad acts on grad φ, which is a vector field, but we know that grad onlyacts on scalar fields (although in fact we will see in chapter 26 that we can <strong>for</strong>mthe outer product of the del operator with a vector to give a tensor, but that neednot concern us here).Of the five valid combinations of grad, div <strong>and</strong> curl, two are identically zero,namelycurl grad φ = ∇×∇φ = 0, (10.37)div curl a = ∇ · (∇×a) =0. (10.38)From (10.37), we see that if a is derived from the gradient of some scalar functionsuch that a = ∇φ then it is necessarily irrotational (∇ ×a = 0). We also notethat if a is an irrotational vector field then another irrotational vector field isa + ∇φ + c, whereφ is any scalar field <strong>and</strong> c is a constant vector. This followssince∇×(a + ∇φ + c) =∇×a + ∇×∇φ = 0.Similarly, from (10.38) we may infer that if b is the curl of some vector field asuch that b = ∇×a then b is solenoidal (∇ · b = 0). Obviously, if b is solenoidal<strong>and</strong> c is any constant vector then b + c is also solenoidal.The three remaining combinations of grad, div <strong>and</strong> curl arediv grad φ = ∇ · ∇φ = ∇ 2 φ = ∂2 φ∂x 2 + ∂2 φ∂y 2 + ∂2 φ∂z 2 , (10.39)grad div a = ∇(∇ · a),( ∂ 2 a x=∂x 2++ ∂2 a y∂x∂y + ∂2 a z∂x∂z) ( ∂ 2 )a xi +∂y∂x + ∂2 a y∂y 2 + ∂2 a zj∂y∂z( ∂ 2 )a x∂z∂x + ∂2 a y∂z∂y + ∂2 a z∂z 2 k, (10.40)curl curl a = ∇×(∇×a) =∇(∇ · a) −∇ 2 a, (10.41)where (10.39) <strong>and</strong> (10.40) are expressed in Cartesian coordinates. In (10.41), theterm ∇ 2 a has the linear differential operator ∇ 2 acting on a vector (as opposed toa scalar as in (10.39)), which of course consists of a sum of unit vectors multipliedby components. Two cases arise.(i) If the unit vectors are constants (i.e. they are independent of the values ofthe coordinates) then the differential operator gives a non-zero contributiononly when acting upon the components, the unit vectors being merelymultipliers.356


10.9 CYLINDRICAL AND SPHERICAL POLAR COORDINATES(ii) If the unit vectors vary as the values of the coordinates change (i.e. arenot constant in direction throughout the whole space) then the derivativesof these vectors appear as contributions to ∇ 2 a.Cartesian coordinates are an example of the first case in which each componentsatisfies (∇ 2 a) i = ∇ 2 a i . In this case (10.41) can be applied to each componentseparately:[∇×(∇×a)] i =[∇(∇ · a)] i −∇ 2 a i . (10.42)However, cylindrical <strong>and</strong> spherical polar coordinates come in the second class.For them (10.41) is still true, but the further step to (10.42) cannot be made.More complicated vector operator relations may be proved using the relationsgiven above.◮Show thatwhere φ <strong>and</strong> ψ are scalar fields.∇ · (∇φ ×∇ψ) =0,From the previous section we have∇ · (a × b) =b · (∇×a) − a · (∇×b).If we let a = ∇φ <strong>and</strong> b = ∇ψ then we obtain∇ · (∇φ ×∇ψ) =∇ψ · (∇×∇φ) −∇φ · (∇×∇ψ) =0, (10.43)since ∇×∇φ =0=∇×∇ψ, from (10.37). ◭10.9 Cylindrical <strong>and</strong> spherical polar coordinatesThe operators we have discussed in this chapter, i.e. grad, div, curl <strong>and</strong> ∇ 2 ,have all been defined in terms of Cartesian coordinates, but <strong>for</strong> many physicalsituations other coordinate systems are more natural. For example, many systems,such as an isolated charge in space, have spherical symmetry <strong>and</strong> spherical polarcoordinates would be the obvious choice. For axisymmetric systems, such as fluidflow in a pipe, cylindrical polar coordinates are the natural choice. The physicallaws governing the behaviour of the systems are often expressed in terms ofthe vector operators we have been discussing, <strong>and</strong> so it is necessary to be ableto express these operators in these other, non-Cartesian, coordinates. We firstconsider the two most common non-Cartesian coordinate systems, i.e. cylindrical<strong>and</strong> spherical polars, <strong>and</strong> go on to discuss general curvilinear coordinates in thenext section.10.9.1 Cylindrical polar coordinatesAs shown in figure 10.7, the position of a point in space P having Cartesiancoordinates x, y, z may be expressed in terms of cylindrical polar coordinates357


VECTOR CALCULUSρ, φ, z, wherex = ρ cos φ, y = ρ sin φ, z = z, (10.44)<strong>and</strong> ρ ≥ 0, 0 ≤ φ


10.9 CYLINDRICAL AND SPHERICAL POLAR COORDINATESzê zPê φê ρrkziOjρyφxFigure 10.7 Cylindrical polar coordinates ρ, φ, z.zρdφdzdρφρdφρdφyxFigure 10.8 The element of volume in cylindrical polar coordinates is givenby ρdρdφdz.Factors, such as the ρ in ρdφ, that multiply the coordinate differentials to givedistances are known as scale factors. From (10.52), the scale factors <strong>for</strong> the ρ-, φ-<strong>and</strong> z- coordinates are there<strong>for</strong>e 1, ρ <strong>and</strong> 1 respectively.The magnitude ds of the displacement dr is given in cylindrical polar coordinatesby(ds) 2 = dr · dr =(dρ) 2 + ρ 2 (dφ) 2 +(dz) 2 ,where in the second equality we have used the fact that the basis vectors areorthonormal. We can also find the volume element in a cylindrical polar system(see figure 10.8) by calculating the volume of the infinitesimal parallelepiped359


VECTOR CALCULUS∇Φ = ∂Φ∂ρ êρ + 1 ∂Φρ ∂φ êφ + ∂Φ∂z êz∇ · a = 1 ∂ρ ∂ρ (ρa ρ)+ 1 ∂a φρ ∂φ + ∂a z∂z∣ ∇×a = 1 ê ρ ρê φ ê z ∣∣∣∣∣∣∣∂ ∂ ∂ρ∂ρ ∂φ ∂z∣ a ρ ρa φ a z∇ 2 Φ = 1 (∂ρ ∂Φ )+ 1 ∂ 2 Φρ ∂ρ ∂ρ ρ 2 ∂φ + ∂2 Φ2 ∂z 2Table 10.2 Vector operators in cylindrical polar coordinates; Φ is a scalarfield <strong>and</strong> a is a vector field.defined by the vectors dρ ê ρ , ρdφê φ <strong>and</strong> dz ê z :dV = |dρ ê ρ · (ρdφê φ × dz ê z )| = ρdρdφdz,which again uses the fact that the basis vectors are orthonormal. For a simplecoordinate system such as cylindrical polars the expressions <strong>for</strong> (ds) 2 <strong>and</strong> dV areobvious from the geometry.We will now express the vector operators discussed in this chapter in termsof cylindrical polar coordinates. Let us consider a vector field a(ρ, φ, z) <strong>and</strong>ascalar field Φ(ρ, φ, z), where we use Φ <strong>for</strong> the scalar field to avoid confusion withthe azimuthal angle φ. We must first write the vector field in terms of the basisvectors of the cylindrical polar coordinate system, i.e.a = a ρ ê ρ + a φ ê φ + a z ê z ,where a ρ , a φ <strong>and</strong> a z are the components of a in the ρ-, φ- <strong>and</strong>z- directionsrespectively. The expressions <strong>for</strong> grad, div, curl <strong>and</strong> ∇ 2 can then be calculated<strong>and</strong> are given in table 10.2. Since the derivations of these expressions are rathercomplicated we leave them until our discussion of general curvilinear coordinatesin the next section; the reader could well postpone examination of these <strong>for</strong>malproofs until some experience of using the expressions has been gained.◮Express the vector field a = yz i − y j + xz 2 k in cylindrical polar coordinates, <strong>and</strong> hencecalculate its divergence. Show that the same result is obtained by evaluating the divergencein Cartesian coordinates.The basis vectors of the cylindrical polar coordinate system are given in (10.49)–(10.51).Solving these equations simultaneously <strong>for</strong> i, j <strong>and</strong> k we obtaini =cosφ ê ρ − sin φ ê φj =sinφ ê ρ +cosφ ê φk = ê z .360


10.9 CYLINDRICAL AND SPHERICAL POLAR COORDINATESzê rê φPê θkθriOjyφxFigure 10.9 Spherical polar coordinates r, θ, φ.Substituting these relations <strong>and</strong> (10.44) into the expression <strong>for</strong> a we finda = zρ sin φ (cos φ ê ρ − sin φ ê φ ) − ρ sin φ (sin φ ê ρ +cosφ ê φ )+z 2 ρ cos φ ê z=(zρsin φ cos φ − ρ sin 2 φ) ê ρ − (zρsin 2 φ + ρ sin φ cos φ) ê φ + z 2 ρ cos φ ê z .Substituting into the expression <strong>for</strong> ∇ · a given in table 10.2,∇ · a =2z sin φ cos φ − 2sin 2 φ − 2z sin φ cos φ − cos 2 φ +sin 2 φ +2zρcos φ=2zρcos φ − 1.Alternatively, <strong>and</strong> much more quickly in this case, we can calculate the divergencedirectly in Cartesian coordinates. We obtain∇ · a = ∂a x∂x + ∂a y∂y + ∂a z=2zx − 1,∂zwhich on substituting x = ρ cos φ yields the same result as the calculation in cylindricalpolars. ◭Finally, we note that similar results can be obtained <strong>for</strong> (two-dimensional)polar coordinates in a plane by omitting the z-dependence. For example, (ds) 2 =(dρ) 2 + ρ 2 (dφ) 2 , while the element of volume is replaced by the element of areadA = ρdρdφ.10.9.2 Spherical polar coordinatesAs shown in figure 10.9, the position of a point in space P , with Cartesiancoordinates x, y, z, may be expressed in terms of spherical polar coordinatesr, θ, φ, wherex = r sin θ cos φ, y = r sin θ sin φ, z = r cos θ, (10.53)361


VECTOR CALCULUS<strong>and</strong> r ≥ 0, 0 ≤ θ ≤ π <strong>and</strong> 0 ≤ φ


10.9 CYLINDRICAL AND SPHERICAL POLAR COORDINATES∇Φ = ∂Φ∂r êr + 1 ∂Φr ∂θ êθ + 1 ∂Φr sin θ ∂φ êφ∇ · a =∇×a =∇ 2 Φ =1 ∂r 2 ∂r (r2 a r )+ 1 ∂r sin θ ∂θ (sin θa θ)+ 1 ∂a φr sin θ ∂φ∣ ê r rê θ r sin θ ê φ ∣∣∣∣∣∣∣1∂ ∂ ∂r 2 sin θ∂r ∂θ ∂φ∣ a r ra θ r sin θa φ(1 ∂r 2 ∂Φ )+ 1 (∂sin θ ∂Φ )+r 2 ∂r ∂r r 2 sin θ ∂θ ∂θ1r 2 sin 2 θ∂ 2 Φ∂φ 2Table 10.3 Vector operators in spherical polar coordinates; Φ is a scalar field<strong>and</strong> a is a vector field.zdφdrθrdθrdθr sin θdφφdφyr sin θr sin θdφxFigure 10.10 The element of volume in spherical polar coordinates is givenby r 2 sin θdrdθdφ.we can rewrite the first term on the RHS as follows:(1 ∂r 2 r 2 ∂Φ )= 1 ∂ 2∂r ∂r r ∂r 2 (rΦ),which can often be useful in shortening calculations.363


VECTOR CALCULUS10.10 General curvilinear coordinatesAs indicated earlier, the contents of this section are more <strong>for</strong>mal <strong>and</strong> technicallycomplicated than hitherto. The section could be omitted until the reader has hadsome experience of using its results.Cylindrical <strong>and</strong> spherical polars are just two examples of what are calledgeneral curvilinear coordinates. In the general case, the position of a point Phaving Cartesian coordinates x, y, z may be expressed in terms of the threecurvilinear coordinates u 1 ,u 2 ,u 3 ,where<strong>and</strong> similarlyx = x(u 1 ,u 2 ,u 3 ), y = y(u 1 ,u 2 ,u 3 ), z = z(u 1 ,u 2 ,u 3 ),u 1 = u 1 (x, y, z), u 2 = u 2 (x, y, z), u 3 = u 3 (x, y, z).We assume that all these functions are continuous, differentiable <strong>and</strong> have asingle-valued inverse, except perhaps at or on certain isolated points or lines,so that there is a one-to-one correspondence between the x, y, z <strong>and</strong> u 1 ,u 2 ,u 3systems. The u 1 -, u 2 -<strong>and</strong>u 3 - coordinate curves of a general curvilinear systemare analogous to the x-, y- <strong>and</strong>z- axes of Cartesian coordinates. The surfacesu 1 = c 1 , u 2 = c 2 <strong>and</strong> u 3 = c 3 , where c 1 ,c 2 ,c 3 are constants, are called thecoordinate surfaces <strong>and</strong> each pair of these surfaces has its intersection in a curvecalled a coordinate curve or line (see figure 10.11).If at each point in space the three coordinate surfaces passing through the pointmeet at right angles then the curvilinear coordinate system is called orthogonal.For example, in spherical polars u 1 = r, u 2 = θ, u 3 = φ <strong>and</strong> the three coordinatesurfaces passing through the point (R,Θ, Φ) are the sphere r = R, the circularcone θ = Θ <strong>and</strong> the plane φ = Φ, which intersect at right angles at thatpoint. There<strong>for</strong>e spherical polars <strong>for</strong>m an orthogonal coordinate system (as docylindrical polars) .If r(u 1 ,u 2 ,u 3 ) is the position vector of the point P then e 1 = ∂r/∂u 1 is a vectortangent to the u 1 -curve at P (<strong>for</strong> which u 2 <strong>and</strong> u 3 are constants) in the directionof increasing u 1 . Similarly, e 2 = ∂r/∂u 2 <strong>and</strong> e 3 = ∂r/∂u 3 are vectors tangent tothe u 2 -<strong>and</strong>u 3 - curves at P in the direction of increasing u 2 <strong>and</strong> u 3 respectively.Denoting the lengths of these vectors by h 1 , h 2 <strong>and</strong> h 3 ,theunit vectors in each ofthese directions are given byê 1 = 1 ∂r, ê 2 = 1 ∂r, ê 3 = 1 ∂r,h 1 ∂u 1 h 2 ∂u 2 h 3 ∂u 3where h 1 = |∂r/∂u 1 |, h 2 = |∂r/∂u 2 | <strong>and</strong> h 3 = |∂r/∂u 3 |.The quantities h 1 , h 2 , h 3 are the scale factors of the curvilinear coordinatesystem. The element of distance associated with an infinitesimal change du i inone of the coordinates is h i du i . In the previous section we found that the scale364


10.10 GENERAL CURVILINEAR COORDINATESuzê 3 ˆɛ 3u 1 = c 1u 2 = c 2 ˆɛê 21 u 2P ê 2u 13ˆɛ 1ku 3 = c 3iOjyxFigure 10.11General curvilinear coordinates.factors <strong>for</strong> cylindrical <strong>and</strong> spherical polar coordinates were<strong>for</strong> cylindrical polars h ρ =1, h φ = ρ, h z =1,<strong>for</strong> spherical polars h r =1, h θ = r, h φ = r sin θ.Although the vectors e 1 , e 2 , e 3 <strong>for</strong>m a perfectly good basis <strong>for</strong> the curvilinearcoordinate system, it is usual to work with the corresponding unit vectors ê 1 , ê 2 ,ê 3 . For an orthogonal curvilinear coordinate system these unit vectors <strong>for</strong>m anorthonormal basis.An infinitesimal vector displacement in general curvilinear coordinates is givenby, from (10.19),dr = ∂r du 1 + ∂r du 2 + ∂r du 3 (10.55)∂u 1 ∂u 2 ∂u 3= du 1 e 1 + du 2 e 2 + du 3 e 3 (10.56)= h 1 du 1 ê 1 + h 2 du 2 ê 2 + h 3 du 3 ê 3 . (10.57)In the case of orthogonal curvilinear coordinates, where the ê iperpendicular, the element of arc length is given byare mutually(ds) 2 = dr · dr = h 2 1(du 1 ) 2 + h 2 2(du 2 ) 2 + h 2 3(du 3 ) 2 . (10.58)The volume element <strong>for</strong> the coordinate system is the volume of the infinitesimalparallelepiped defined by the vectors (∂r/∂u i ) du i = du i e i = h i du i ê i ,<strong>for</strong>i =1, 2, 3.365


VECTOR CALCULUSFor orthogonal coordinates this is given bydV = |du 1 e 1 · (du 2 e 2 × du 3 e 3 )|= |h 1 ê 1 · (h 2 ê 2 × h 3 ê 3 )| du 1 du 2 du 3= h 1 h 2 h 3 du 1 du 2 du 3 .Now, in addition to the set {ê i }, i =1, 2, 3, there exists another useful set ofthree unit basis vectors at P .Since∇u 1 is a vector normal to the surface u 1 = c 1 ,a unit vector in this direction is ˆɛ 1 = ∇u 1 /|∇u 1 |. Similarly, ˆɛ 2 = ∇u 2 /|∇u 2 | <strong>and</strong>ˆɛ 3 = ∇u 3 /|∇u 3 | are unit vectors normal to the surfaces u 2 = c 2 <strong>and</strong> u 3 = c 3respectively.There<strong>for</strong>e at each point P in a curvilinear coordinate system, there exist, ingeneral, two sets of unit vectors: {ê i }, tangent to the coordinate curves, <strong>and</strong> {ˆɛ i },normal to the coordinate surfaces. A vector a can be written in terms of eitherset of unit vectors:a = a 1 ê 1 + a 2 ê 2 + a 3 ê 3 = A 1ˆɛ 1 + A 2ˆɛ 2 + A 3ˆɛ 3 ,where a 1 , a 2 , a 3 <strong>and</strong> A 1 , A 2 , A 3 are the components of a in the two systems. Itmay be shown that the two bases become identical if the coordinate system isorthogonal.Instead of the unit vectors discussed above, we could instead work directly withthe two sets of vectors {e i = ∂r/∂u i } <strong>and</strong> {ɛ i = ∇u i },whicharenot,ingeneral,ofunit length. We can then write a vector a asa = α 1 e 1 + α 2 e 2 + α 3 e 3 = β 1 ɛ 1 + β 2 ɛ 2 + β 3 ɛ 3 ,or more explicitly as∂r ∂r ∂ra = α 1 + α 2 + α 3 = β 1 ∇u 1 + β 2 ∇u 2 + β 3 ∇u 3 ,∂u 1 ∂u 2 ∂u 3where α 1 ,α 2 ,α 3 <strong>and</strong> β 1 ,β 2 ,β 3 are called the contravariant <strong>and</strong> covariant componentsof a respectively. A more detailed discussion of these components, inthe context of tensor analysis, is given in chapter 26. The (in general) non-unitbases {e i } <strong>and</strong> {ɛ i } are often the most natural bases in which to express vectorquantities.◮Show that {e i } <strong>and</strong> {ɛ i } are reciprocal systems of vectors.Let us consider the scalar product e i · ɛ j ; using the Cartesian expressions <strong>for</strong> r <strong>and</strong> ∇, weobtaine i · ɛ j = ∂r · ∇u j∂u i)=( ∂x∂u ii + ∂y∂u ij + ∂z∂u ik= ∂x∂u i∂u j∂x + ∂y∂u i∂u j∂y + ∂z∂u i∂u j∂z = ∂u j∂u i.366( ∂uj·∂x i + ∂u j∂y j + ∂u )j∂z k


10.10 GENERAL CURVILINEAR COORDINATESIn the last step we have used the chain rule <strong>for</strong> partial differentiation. There<strong>for</strong>e e i · ɛ j =1if i = j, <strong>and</strong>e i · ɛ j = 0 otherwise. Hence {e i } <strong>and</strong> {ɛ j } are reciprocal systems of vectors. ◭We now derive expressions <strong>for</strong> the st<strong>and</strong>ard vector operators in orthogonalcurvilinear coordinates. Despite the useful properties of the non-unit bases discussedabove, the remainder of our discussion in this section will be in terms ofthe unit basis vectors {ê i }. The expressions <strong>for</strong> the vector operators in cylindrical<strong>and</strong> spherical polar coordinates given in tables 10.2 <strong>and</strong> 10.3 respectively can befound from those derived below by inserting the appropriate scale factors.GradientThe change dΦ in a scalar field Φ resulting from changes du 1 ,du 2 ,du 3 in thecoordinates u 1 ,u 2 ,u 3 is given by, from (5.5),dΦ = ∂Φ du 1 + ∂Φ du 2 + ∂Φ du 3 .∂u 1 ∂u 2 ∂u 3For orthogonal curvilinear coordinates u 1 ,u 2 ,u 3 we find from (10.57), <strong>and</strong> comparisonwith (10.27), that we can write this aswhere ∇Φ is given bydΦ =∇Φ · dr, (10.59)∇Φ = 1 ∂Φê 1 + 1 ∂Φê 2 + 1 ∂Φê 3 . (10.60)h 1 ∂u 1 h 2 ∂u 2 h 3 ∂u 3This implies that the del operator can be written∇ = ê1h 1∂∂u 1+ ê2h 2∂∂u 2+ ê3h 3∂∂u 3.◮Show that <strong>for</strong> orthogonal curvilinear coordinates ∇u i = ê i /h i . Hence show that the twosets of vectors {ê i } <strong>and</strong> {ˆɛ i } are identical in this case.Letting Φ = u i in (10.60) we find immediately that ∇u i = ê i /h i . There<strong>for</strong>e |∇u i | =1/h i ,<strong>and</strong>so ˆɛ i = ∇u i /|∇u i | = h i ∇u i = ê i . ◭DivergenceIn order to derive the expression <strong>for</strong> the divergence of a vector field in orthogonalcurvilinear coordinates, we must first write the vector field in terms of the basisvectors of the coordinate system:The divergence is then given by∇ · a = 1h 1 h 2 h 3[ ∂∂u 1(h 2 h 3 a 1 )+a = a 1 ê 1 + a 2 ê 2 + a 3 ê 3 .∂∂u 2(h 3 h 1 a 2 )+367∂ ](h 1 h 2 a 3 ) .∂u 3(10.61)


VECTOR CALCULUS◮Prove the expression <strong>for</strong> ∇ · a in orthogonal curvilinear coordinates.Let us consider the sub-expression ∇ · (a 1 ê 1 ). Now ê 1 = ê 2 × ê 3 = h 2 ∇u 2 × h 3 ∇u 3 . There<strong>for</strong>e∇ · (a 1 ê 1 )=∇ · (a 1 h 2 h 3 ∇u 2 ×∇u 3 ),= ∇(a 1 h 2 h 3 ) · (∇u 2 ×∇u 3 )+a 1 h 2 h 3 ∇ · (∇u 2 ×∇u 3 ).However, ∇ · (∇u 2 ×∇u 3 ) = 0, from (10.43), so we obtain( )ê2∇ · (a 1 ê 1 )=∇(a 1 h 2 h 3 ) · × ê3ê 1= ∇(a 1 h 2 h 3 ) · ;h 2 h 3 h 2 h 3letting Φ = a 1 h 2 h 3 in (10.60) <strong>and</strong> substituting into the above equation, we find∇ · (a 1 ê 1 )= 1 ∂(a 1 h 2 h 3 ).h 1 h 2 h 3 ∂u 1Repeating the analysis <strong>for</strong> ∇ · (a 2 ê 2 )<strong>and</strong>∇ · (a 3 ê 3 ), <strong>and</strong> adding the results we obtain (10.61),as required. ◭LaplacianIn the expression <strong>for</strong> the divergence (10.61), leta = ∇Φ = 1 ∂Φê 1 + 1 ∂Φê 2 + 1 ∂Φê 3 ,h 1 ∂u 1 h 2 ∂u 2 h 3 ∂u 3where we have used (10.60). We then obtain∇ 2 Φ= 1 [ ( )∂ h2 h 3 ∂Φ+ ∂ ( )h3 h 1 ∂Φ+ ∂ ( )]h1 h 2 ∂Φ,h 1 h 2 h 3 ∂u 1 h 1 ∂u 1 ∂u 2 h 2 ∂u 2 ∂u 3 h 3 ∂u 3which is the expression <strong>for</strong> the Laplacian in orthogonal curvilinear coordinates.CurlThe curl of a vector field a = a 1 ê 1 + a 2 ê 2 + a 3 ê 3coordinates is given byin orthogonal curvilinear∣ ∣ ∣∣∣∣∣∣∣∣ h 1 ê 1 h 2 ê 2 h 3 ê 3 ∣∣∣∣∣∣∣∣∇×a = 1 ∂ ∂ ∂. (10.62)h 1 h 2 h 3 ∂u 1 ∂u 2 ∂u 3h 1 a 1 h 2 a 2 h 3 a 3◮Prove the expression <strong>for</strong> ∇×a in orthogonal curvilinear coordinates.Let us consider the sub-expression ∇×(a 1 ê 1 ). Since ê 1 = h 1 ∇u 1 we have∇×(a 1 ê 1 )=∇×(a 1 h 1 ∇u 1 ),= ∇(a 1 h 1 ) ×∇u 1 + a 1 h 1 ∇×∇u 1 .But ∇×∇u 1 =0,soweobtain∇×(a 1 ê 1 )=∇(a 1 h 1 ) × ê1h 1.368


10.11 EXERCISES∇Φ =∇ · a =∇×a =∇ 2 Φ =1 ∂Φê 1 + 1 ∂Φê 2 + 1 ∂Φê 3h 1 ∂u 1 h 2 ∂u 2 h 3 ∂u 3[1 ∂(h 2 h 3 a 1 )+h 1 h 2 h 3 ∂u 1∣ ∣ ∣∣∣∣∣∣ h 1 ê 1 h 2 ê 2 h 3 ê 3 ∣∣∣∣∣∣1 ∂ ∂ ∂h 1 h 2 h 3 ∂u 1 ∂u 2 ∂u 3h 1 a 1 h 2 a 2 h 3 a 3∂∂u 2(h 3 h 1 a 2 )+∂ ](h 1 h 2 a 3 )∂u 3[ ( )1 ∂ h2 h 3 ∂Φ+ ∂ ( )h3 h 1 ∂Φ+ ∂ ( )]h1 h 2 ∂Φh 1 h 2 h 3 ∂u 1 h 1 ∂u 1 ∂u 2 h 2 ∂u 2 ∂u 3 h 3 ∂u 3Table 10.4 Vector operators in orthogonal curvilinear coordinates u 1 ,u 2 ,u 3 .Φ is a scalar field <strong>and</strong> a is a vector field.Letting Φ = a 1 h 1 in (10.60) <strong>and</strong> substituting into the above equation, we find∇×(a 1 ê 1 )=ê2 ∂(a 1 h 1 ) −ê3 ∂(a 1 h 1 ).h 3 h 1 ∂u 3 h 1 h 2 ∂u 2The corresponding analysis of ∇×(a 2 ê 2 ) produces terms in ê 3 <strong>and</strong> ê 1 , whilst that of∇×(a 3 ê 3 ) produces terms in ê 1 <strong>and</strong> ê 2 . When the three results are added together, thecoefficients multiplying ê 1 , ê 2 <strong>and</strong> ê 3 are the same as those obtained by writing out (10.62)explicitly, thus proving the stated result. ◭The general expressions <strong>for</strong> the vector operators in orthogonal curvilinearcoordinates are shown <strong>for</strong> reference in table 10.4. The explicit results <strong>for</strong> cylindrical<strong>and</strong> spherical polar coordinates, given in tables 10.2 <strong>and</strong> 10.3 respectively, areobtained by substituting the appropriate set of scale factors in each case.A discussion of the expressions <strong>for</strong> vector operators in tensor <strong>for</strong>m, whichare valid even <strong>for</strong> non-orthogonal curvilinear coordinate systems, is given inchapter 26.10.11 Exercises10.1 Evaluate the integral∫ [a(ḃ · a + b · ȧ)+ȧ(b · a) − 2(ȧ · a)b − ḃ|a|2] dtin which ȧ, ḃ are the derivatives of a, b with respect to t.10.2 At time t = 0, the vectors E <strong>and</strong> B are given by E = E 0 <strong>and</strong> B = B 0 ,wheretheunit vectors, E 0 <strong>and</strong> B 0 are fixed <strong>and</strong> orthogonal. The equations of motion aredEdt = E 0 + B × E 0 ,dBdt = B 0 + E × B 0 .Find E <strong>and</strong> B at a general time t, showing that after a long time the directionsof E <strong>and</strong> B have almost interchanged.369


VECTOR CALCULUS10.3 The general equation of motion of a (non-relativistic) particle of mass m <strong>and</strong>charge q when it is placed in a region where there is a magnetic field B <strong>and</strong> anelectric field E ism¨r = q(E + ṙ × B);here r is the position of the particle at time t <strong>and</strong> ṙ = dr/dt, etc. Write this asthree separate equations in terms of the Cartesian components of the vectorsinvolved.For the simple case of crossed uni<strong>for</strong>m fields E = Ei, B = Bj, inwhichtheparticle starts from the origin at t =0withṙ = v 0 k, find the equations of motion<strong>and</strong> show the following:(a) if v 0 = E/B then the particle continues its initial motion;(b) if v 0 = 0 then the particle follows the space curve given in terms of theparameter ξ byx = mEmE(1 − cos ξ), y =0, z = (ξ − sin ξ).B 2 q B 2 qInterpret this curve geometrically <strong>and</strong> relate ξ to t. Show that the totaldistance travelled by the particle after time t is given by∫2E tBqt′B ∣sin 0 2m ∣ dt′ .10.4 Use vector methods to find the maximum angle to the horizontal at which a stonemay be thrown so as to ensure that it is always moving away from the thrower.10.5 If two systems of coordinates with a common origin O are rotating with respectto each other, the measured accelerations differ in the two systems. Denotingby r <strong>and</strong> r ′ position vectors in frames OXY Z <strong>and</strong> OX ′ Y ′ Z ′ , respectively, theconnection between the two is¨r ′ = ¨r + ˙ω × r +2ω × ṙ + ω × (ω × r),where ω is the angular velocity vector of the rotation of OXY Z with respect toOX ′ Y ′ Z ′ (taken as fixed). The third term on the RHS is known as the Coriolisacceleration, whilst the final term gives rise to a centrifugal <strong>for</strong>ce.Consider the application of this result to the firing of a shell of mass m froma stationary ship on the steadily rotating earth, working to the first order inω (= 7.3 × 10 −5 rad s −1 ). If the shell is fired with velocity v at time t =0<strong>and</strong>onlyreaches a height that is small compared with the radius of the earth, show thatits acceleration, as recorded on the ship, is given approximately by¨r = g − 2ω × (v + gt),where mg is the weight of the shell measured on the ship’s deck.The shell is fired at another stationary ship (a distance s away) <strong>and</strong> v is suchthat the shell would have hit its target had there been no Coriolis effect.(a) Show that without the Coriolis effect the time of flight of the shell wouldhave been τ = −2g · v/g 2 .(b) Show further that when the shell actually hits the sea it is off-target byapproximately2τg [(g × ω) · v](gτ + v) − (ω × 2 v)τ2 − 1 3 (ω × g)τ3 .(c) Estimate the order of magnitude ∆ of this miss <strong>for</strong> a shell <strong>for</strong> which theinitial speed v is 300 m s −1 , firing close to its maximum range (v makes anangle of π/4 with the vertical) in a northerly direction, whilst the ship isstationed at latitude 45 ◦ North.370


10.11 EXERCISES10.6 Prove that <strong>for</strong> a space curve r = r(s), where s is the arc length measured alongthe curve from a fixed point, the triple scalar product( ) drds × d2 r· d3 rds 2 ds 3at any point on the curve has the value κ 2 τ,whereκ is the curvature <strong>and</strong> τ thetorsion at that point.10.7 For the twisted space curve y 3 +27axz − 81a 2 y = 0, given parametrically byx = au(3 − u 2 ), y =3au 2 , z = au(3 + u 2 ),show that the following hold:(a) ds/du =3 √ 2a(1 + u 2 ), where s is the distance along the curve measured fromthe origin;(b) the length of the curve from the origin to the Cartesian point (2a, 3a, 4a) is4 √ 2a;(c) the radius of curvature at the point with parameter u is 3a(1 + u 2 ) 2 ;(d) the torsion τ <strong>and</strong> curvature κ at a general point are equal;(e) any of the Frenet–Serret <strong>for</strong>mulae that you have not already used directlyare satisfied.10.8 The shape of the curving slip road joining two motorways, that cross at rightangles <strong>and</strong> are at vertical heights z =0<strong>and</strong>z = h, can be approximated by thespace curve√2hr =π( zπ)√2hln cos i +2h π( zπ)ln sin j + zk.2hShow that the radius of curvature ρ of the slip road is (2h/π)cosec (zπ/h) atheight z <strong>and</strong> that the torsion τ = −1/ρ. To shorten the algebra, set z =2hθ/π<strong>and</strong> use θ as the parameter.10.9 In a magnetic field, field lines are curves to which the magnetic induction B iseverywhere tangential. By evaluating dB/ds, wheres is the distance measuredalong a field line, prove that the radius of curvature at any point on a line isgiven byB 3ρ =|B × (B · ∇)B| .10.10 Find the areas of the given surfaces using parametric coordinates.(a) Using the parameterisation x = u cos φ, y = u sin φ, z = u cot Ω, find thesloping surface area of a right circular cone of semi-angle Ω whose base hasradius a. Verifythatitisequalto 1 ×perimeter of the base ×slope height.2(b) Using the same parameterization as in (a) <strong>for</strong> x <strong>and</strong> y, <strong>and</strong> an appropriatechoice <strong>for</strong> z, find the surface area between the planes z =0<strong>and</strong>z = Z ofthe paraboloid of revolution z = α(x 2 + y 2 ).10.11 Parameterising the hyperboloidx 2a + y22 b − z22 c =1 2by x = a cos θ sec φ, y = b sin θ sec φ, z = c tan φ, show that an area element onits surface isdS =sec 2 φ [ c 2 sec 2 φ ( b 2 cos 2 θ + a 2 sin 2 θ ) + a 2 b 2 tan 2 φ ] 1/2dθ dφ.371


VECTOR CALCULUSUse this <strong>for</strong>mula to show that the area of the curved surface x 2 + y 2 − z 2 = a 2between the planes z =0<strong>and</strong>z =2a is(πa 2 6+ √ 1 sinh −1 2 √ )2 .210.12 For the functionz(x, y) =(x 2 − y 2 )e −x2 −y 2 ,find the location(s) at which the steepest gradient occurs. What are the magnitude<strong>and</strong> direction of that gradient? The algebra involved is easier if plane polarcoordinates are used.10.13 Verify by direct calculation that∇ · (a × b) =b · (∇×a) − a · (∇×b).10.14 In the following exercises, a, b <strong>and</strong> c are vector fields.(a) Simplify∇×a(∇ · a) + a × [∇×(∇×a)] + a ×∇ 2 a.(b) By explicitly writing out the terms in Cartesian coordinates, prove that[c · (b · ∇) − b · (c · ∇)] a =(∇×a) · (b × c).(c) Prove that a × (∇×a) =∇( 1 2 a2 ) − (a · ∇)a.10.15 Evaluate the Laplacian of the functionzx 2ψ(x, y, z) =x 2 + y 2 + z 2(a) directly in Cartesian coordinates, <strong>and</strong> (b) after changing to a spherical polarcoordinate system. Verify that, as they must, the two methods give the sameresult.10.16 Verify that (10.42) is valid <strong>for</strong> each component separately when a is the Cartesianvector x 2 y i + xyz j + z 2 y k, by showing that each side of the equation is equal toz i +(2x +2z) j + x k.10.17 The (Maxwell) relationship between a time-independent magnetic field B <strong>and</strong> thecurrent density J (measured in SI units in A m −2 ) producing it,∇×B = µ 0 J,can be applied to a long cylinder of conducting ionised gas which, in cylindricalpolar coordinates, occupies the region ρa<strong>and</strong> B = B(ρ) <strong>for</strong>ρ


10.11 EXERCISES(a) For cylindrical polar coordinates ρ, φ, z, evaluate the derivatives of the threeunit vectors with respect to each of the coordinates, showing that only ∂ê ρ /∂φ<strong>and</strong> ∂ê φ /∂φ are non-zero.(i) Hence evaluate ∇ 2 a when a is the vector ê ρ , i.e. a vector of unit magnitudeeverywhere directed radially outwards <strong>and</strong> expressed by a ρ =1,a φ =a z =0.(ii) Note that it is trivially obvious that ∇×a = 0 <strong>and</strong> hence that equation(10.41) requires that ∇(∇ · a) =∇ 2 a.(iii) Evaluate ∇(∇ · a) <strong>and</strong> show that the latter equation holds, but that[∇(∇ · a)] ρ ≠ ∇ 2 a ρ .(b) Rework the same problem in Cartesian coordinates (where, as it happens,the algebra is more complicated).10.19 Maxwell’s equations <strong>for</strong> electromagnetism in free space (i.e. in the absence ofcharges, currents <strong>and</strong> dielectric or magnetic media) can be written(i) ∇ · B =0, (ii) ∇ · E =0,(iii) ∇×E + ∂B∂t = 0, (iv) ∇×B − 1 ∂Ec 2 ∂t = 0.A vector A is defined by B = ∇×A, <strong>and</strong>ascalarφ by E = −∇φ − ∂A/∂t. Showthat if the condition(v) ∇ · A + 1 ∂φc 2 ∂t =0is imposed (this is known as choosing the Lorentz gauge), then A <strong>and</strong> φ satisfywave equations as follows:(vi) ∇ 2 φ − 1 ∂ 2 φc 2 ∂t =0, 2(vii) ∇ 2 A − 1 ∂ 2 Ac 2 ∂t = 0. 2The reader is invited to proceed as follows.(a) Verify that the expressions <strong>for</strong> B <strong>and</strong> E in terms of A <strong>and</strong> φ are consistentwith (i) <strong>and</strong> (iii).(b) Substitute <strong>for</strong> E in (ii) <strong>and</strong> use the derivative with respect to time of (v) toeliminate A from the resulting expression. Hence obtain (vi).(c) Substitute <strong>for</strong> B <strong>and</strong> E in (iv) in terms of A <strong>and</strong> φ. Then use the gradient of(v) to simplify the resulting equation <strong>and</strong> so obtain (vii).10.20 In a description of the flow of a very viscous fluid that uses spherical polarcoordinates with axial symmetry, the components of the velocity field u are givenin terms of the stream function ψ by1 ∂ψu r =r 2 sin θ ∂θ , u θ = −1 ∂ψr sin θ ∂r .Find an explicit expression <strong>for</strong> the differential operator E defined byEψ = −(r sin θ)(∇×u) φ .The stream function satisfies the equation of motion E 2 ψ = 0 <strong>and</strong>, <strong>for</strong> the flow ofa fluid past a sphere, takes the <strong>for</strong>m ψ(r, θ) =f(r)sin 2 θ. Show that f(r) satisfiesthe (ordinary) differential equationr 4 f (4) − 4r 2 f ′′ +8rf ′ − 8f =0.373


VECTOR CALCULUS10.21 Paraboloidal coordinates u, v, φ are defined in terms of Cartesian coordinates byx = uv cos φ, y = uv sin φ, z = 1 2 (u2 − v 2 ).Identify the coordinate surfaces in the u, v, φ system. Verify that each coordinatesurface (u = constant, say) intersects every coordinate surface on which one ofthe other two coordinates (v, say) is constant. Show further that the system ofcoordinates is an orthogonal one <strong>and</strong> determine its scale factors. Prove that theu-component of ∇×a is given by(1 aφ(u 2 + v 2 ) 1/2 v + ∂a )φ− 1 ∂a v∂v uv ∂φ .10.22 Non-orthogonal curvilinear coordinates are difficult to work with <strong>and</strong> should beavoided if at all possible, but the following example is provided to illustrate thecontent of section 10.10.In a new coordinate system <strong>for</strong> the region of space in which the Cartesiancoordinate z satisfies z ≥ 0, the position of a point r is given by (α 1 ,α 2 ,R), whereα 1 <strong>and</strong> α 2 are respectively the cosines of the angles made by r with the x- <strong>and</strong>ycoordinateaxes of a Cartesian system <strong>and</strong> R = |r|. The ranges are −1 ≤ α i ≤ 1,0 ≤ R


10.12 HINTS AND ANSWERS(a) Express z <strong>and</strong> the perpendicular distance ρ from P to the z-axis in terms ofu 1 ,u 2 ,u 3 .(b) Evaluate ∂x/∂u i , ∂y/∂u i , ∂z/∂u i ,<strong>for</strong>i =1, 2, 3.(c)Find the Cartesian components of û j <strong>and</strong> hence show that the new coordinatesare mutually orthogonal. Evaluate the scale factors <strong>and</strong> the infinitesimalvolume element in the new coordinate system.(d) Determine <strong>and</strong> sketch the <strong>for</strong>ms of the surfaces u i =constant.(e) Find the most general function f of u 1 only that satisfies ∇ 2 f =0.10.12 Hints <strong>and</strong> answers10.1 Group the term so that they <strong>for</strong>m the total derivatives of compound vectorexpressions. The integral has the value a × (a × b)+h.10.3 For crossed uni<strong>for</strong>m fields, ẍ+(Bq/m) 2 x = q(E −Bv 0 )/m, ÿ =0,mż = qBx+mv 0 ;(b) ξ = Bqt/m; the path is a cycloid in the plane y =0;ds =[(dx/dt) 2 +(dz/dt) 2 ] 1/2 dt.10.5 g = ¨r ′ − ω × (ω × r), where ¨r ′ is the shell’s acceleration measured by an observerfixed in space. To first order in ω, the direction of g is radial, i.e. parallel to ¨r ′ .(a) Note that s is orthogonal to g.(b) If the actual time of flight is T ,use(s +∆)· g =0toshowthatT ≈ τ(1 + 2g −2 (g × ω) · v + ···).In the Coriolis terms, it is sufficient to put T ≈ τ.(c) For this situation (g × ω) · v =0<strong>and</strong>ω × v = 0; τ ≈ 43 s <strong>and</strong> ∆ = 10–15 mto the East.10.7 (a) Evaluate (dr/du) · (dr/du).(b) Integrate the previous result between u =0<strong>and</strong>u =1.(c) ˆt =[ √ 2(1 + u 2 )] −1 [(1 − u 2 )i +2uj +(1+u 2 )k]. Use dˆt/ds =(dˆt/du)/(ds/du);ρ −1 = |dˆt/ds|.(d) ˆn =(1+u 2 ) −1 [−2ui +(1− u 2 )j]. ˆb =[ √ 2(1 + u 2 )] −1 [(u 2 − 1)i − 2uj +(1+u 2 )k].Use dˆb/ds =(dˆb/du)/(ds/du) <strong>and</strong> show that this equals −[3a(1 + u 2 ) 2 ] −1 ˆn.(e) Show that dˆn/ds = τ(ˆb − ˆt) =−2[3 √ 2a(1 + u 2 ) 3 ] −1 [(1 − u 2 )i +2uj].10.9 Note that dB =(dr · ∇)B <strong>and</strong> that B = Bˆt, withˆt = dr/ds. Obtain(B · ∇)B/B =ˆt(dB/ds)+ˆn(B/ρ) <strong>and</strong> then take the vector product of ˆt with this equation.10.11 To integrate sec 2 φ(sec 2 φ +tan 2 φ) 1/2 dφ put tan φ =2 −1/2 sinh ψ.10.13 Work in Cartesian coordinates, regrouping the terms obtained by evaluating thedivergence on the LHS.10.15 (a) 2z(x 2 +y 2 +z 2 ) −3 [(y 2 +z 2 )(y 2 +z 2 −3x 2 )−4x 4 ]; (b) 2r −1 cos θ (1−5sin 2 θ cos 2 φ);both are equal to 2zr −4 (r 2 − 5x 2 ).10.17 Use the <strong>for</strong>mulae given in table 10.2.(a) C = −B 0 /(µ 0 a); B(ρ) =B 0 ρ/a.(b) B 0 ρ 2 /(3a) <strong>for</strong>ρa.(c) [B0 2/(2µ 0)][1 − (ρ/a) 2 ].10.19 Recall that ∇×∇φ = 0 <strong>for</strong> any scalar φ <strong>and</strong> that ∂/∂t <strong>and</strong> ∇ act on differentvariables.10.21 Two sets of paraboloids of revolution about the z-axis <strong>and</strong> the sheaf of planescontaining the z-axis. For constant u, −∞


VECTOR CALCULUS10.23 The tangent vectors are as follows: <strong>for</strong> u = 0, the line joining (1, 0, 0) <strong>and</strong>(−1, 0, 0); <strong>for</strong> v = 0, the line joining (1, 0, 0) <strong>and</strong> (∞, 0, 0); <strong>for</strong> v = π/2, the line(0, 0,z); <strong>for</strong> v = π, the line joining (−1, 0, 0) <strong>and</strong> (−∞, 0, 0).ψ(u) =2tan −1 e u + c, derivedfrom∂[cosh u(∂ψ/∂u)]/∂u =0.376


11Line, surface <strong>and</strong> volume integralsIn the previous chapter we encountered continuously varying scalar <strong>and</strong> vectorfields <strong>and</strong> discussed the action of various differential operators on them. Inaddition to these differential operations, the need often arises to consider theintegration of field quantities along lines, over surfaces <strong>and</strong> throughout volumes.In general the integr<strong>and</strong> may be scalar or vector in nature, but the evaluationof such integrals involves their reduction to one or more scalar integrals, whichare then evaluated. In the case of surface <strong>and</strong> volume integrals this requires theevaluation of double <strong>and</strong> triple integrals (see chapter 6).11.1 Line integralsIn this section we discuss line or path integrals, in which some quantity relatedto the field is integrated between two given points in space, A <strong>and</strong> B, along aprescribed curve C that joins them. In general, we may encounter line integralsof the <strong>for</strong>ms ∫ ∫∫φdr, a · dr, a × dr, (11.1)CCwhere φ is a scalar field <strong>and</strong> a is a vector field. The three integrals themselves arerespectively vector, scalar <strong>and</strong> vector in nature. As we will see below, in physicalapplications line integrals of the second type are by far the most common.The <strong>for</strong>mal definition of a line integral closely follows that of ordinary integrals<strong>and</strong> can be considered as the limit of a sum. We may divide the path C joiningthe points A <strong>and</strong> B into N small line elements ∆r p , p =1,...,N.If(x p ,y p ,z p )isany point on the line element ∆r p then the second type of line integral in (11.1),<strong>for</strong> example, is defined as∫∑Na · dr = lim a(x p ,y p ,z p ) · ∆r p ,CN→∞p=1where it is assumed that all |∆r p |→0asN →∞.377C


LINE, SURFACE AND VOLUME INTEGRALSEach of the line integrals in (11.1) is evaluated over some curve C that may beeither open (A <strong>and</strong> B being distinct points) or closed (the curve C <strong>for</strong>ms a loop,so that A <strong>and</strong> B are coincident). In the case where C is closed, the line integralis written ∮ Cto indicate this. The curve may be given either parametrically byr(u) =x(u)i + y(u)j + z(u)k or by means of simultaneous equations relating x, y, z<strong>for</strong> the given path (in Cartesian coordinates). A full discussion of the differentrepresentations of space curves was given in section 10.3.In general, the value of the line integral depends not only on the end-pointsA <strong>and</strong> B but also on the path C joining them. For a closed curve we must alsospecify the direction around the loop in which the integral is taken. It is usuallytaken to be such that a person walking around the loop C in this directionalways has the region R on his/her left; this is equivalent to traversing C in theanticlockwise direction (as viewed from above).11.1.1 Evaluating line integralsThe method of evaluating a line integral is to reduce it to a set of scalar integrals.It is usual to work in Cartesian coordinates, in which case dr = dx i + dy j + dz k.The first type of line integral in (11.1) then becomes simply∫ ∫∫∫φdr = i φ(x, y, z) dx + j φ(x, y, z) dy + k φ(x, y, z) dz.CCCThe three integrals on the RHS are ordinary scalar integrals that can be evaluatedin the usual way once the path of integration C has been specified. Note that inthe above we have used relations of the <strong>for</strong>m∫ ∫φ i dx = i φdx,which is allowable since the Cartesian unit vectors are of constant magnitude<strong>and</strong> direction <strong>and</strong> hence may be taken out of the integral. If we had been usinga different coordinate system, such as spherical polars, then, as we saw in theprevious chapter, the unit basis vectors would not be constant. In that case thebasis vectors could not be factorised out of the integral.The second <strong>and</strong> third line integrals in (11.1) can also be reduced to a set ofscalar integrals by writing the vector field a in terms of its Cartesian componentsas a = a x i + a y j + a z k,wherea x ,a y ,a z are each (in general) functions of x, y, z.The second line integral in (11.1), <strong>for</strong> example, can then be written as∫ ∫a · dr = (a x i + a y j + a z k) · (dx i + dy j + dz k)CC∫= (a x dx + a y dy + a z dz)C∫ ∫ ∫= a x dx + a y dy + a z dz. (11.2)CC378CC


11.1 LINE INTEGRALSA similar procedure may be followed <strong>for</strong> the third type of line integral in (11.1),which involves a cross product.Line integrals have properties that are analogous to those of ordinary integrals.In particular, the following are useful properties (which we illustrate using thesecond <strong>for</strong>m of line integral in (11.1) but which are valid <strong>for</strong> all three types).(i) Reversing the path of integration changes the sign of the integral. If thepath C along which the line integrals are evaluated has A <strong>and</strong> B as itsend-points then∫ BAa · dr = −∫ ABa · dr.This implies that if the path C is a loop then integrating around the loopin the opposite direction changes the sign of the integral.(ii) If the path of integration is subdivided into smaller segments then the sumof the separate line integrals along each segment is equal to the line integralalong the whole path. So, if P is any point on the path of integration thatlies between the path’s end-points A <strong>and</strong> B then∫ BAa · dr =∫ PAa · dr +∫ BPa · dr.◮Evaluate the line integral I = ∫ a · dr, wherea =(x + y)i +(y − x)j, along each of theCpaths in the xy-plane shown in figure 11.1, namely(i) the parabola y 2 = x from (1, 1) to (4, 2),(ii) the curve x =2u 2 + u +1, y =1+u 2 from (1, 1) to (4, 2),(iii) the line y = 1 from (1, 1) to (4, 1), followed by the line x = 4 from (4, 1)to (4, 2).Since each of the paths lies entirely in the xy-plane, we have dr = dx i + dy j. Wecanthere<strong>for</strong>e write the line integral as∫ ∫I = a · dr = [(x + y) dx +(y − x) dy]. (11.3)CCWe must now evaluate this line integral along each of the prescribed paths.Case (i). Along the parabola y 2 = x we have 2ydy = dx. Substituting <strong>for</strong> x in (11.3)<strong>and</strong> using just the limits on y, weobtainI =∫ (4,2)(1,1)[(x + y) dx +(y − x) dy] =∫ 21[(y 2 + y)2y +(y − y 2 )] dy =11 1 3 .Note that we could just as easily have substituted <strong>for</strong> y <strong>and</strong> obtained an integral in x,which would have given the same result.Case (ii). The second path is given in terms of a parameter u. We could eliminate ubetween the two equations to obtain a relationship between x <strong>and</strong> y directly <strong>and</strong> proceedas above, but it is usually quicker to write the line integral in terms of the parameter u.Along the curve x =2u 2 + u +1, y =1+u 2 we have dx =(4u +1)du <strong>and</strong> dy =2udu.379


LINE, SURFACE AND VOLUME INTEGRALSy(i)(4, 2)(ii)(1, 1)(iii)Figure 11.1 Different possible paths between the points (1, 1) <strong>and</strong> (4, 2).xSubstituting <strong>for</strong> x <strong>and</strong> y in (11.3) <strong>and</strong> writing the correct limits on u, weobtainI ==∫ (4,2)(1,1)∫ 10[(x + y) dx +(y − x) dy][(3u 2 + u + 2)(4u +1)− (u 2 + u)2u] du =10 2 3 .Case (iii). For the third path the line integral must be evaluated along the two linesegments separately <strong>and</strong> the results added together. First, along the line y = 1 we havedy = 0. Substituting this into (11.3) <strong>and</strong> using just the limits on x <strong>for</strong>thissegment,weobtain∫ (4,1)∫ 4[(x + y) dx +(y − x) dy] = (x +1)dx =10 1 . 2(1,1)Next, along the line x = 4 we have dx = 0. Substituting this into (11.3) <strong>and</strong> using just thelimits on y <strong>for</strong> this segment, we obtain∫ (4,2)(4,1)[(x + y) dx +(y − x) dy] =1∫ 21(y − 4) dy = −2 1 2 .The value of the line integral along the whole path is just the sum of the values of the lineintegrals along each segment, <strong>and</strong> is given by I =10 1 2 − 2 1 2 =8.◭When calculating a line integral along some curve C, which is given in termsof x, y <strong>and</strong> z, we are sometimes faced with the problem that the curve C is suchthat x, y <strong>and</strong> z are not single-valued functions of one another over the entirelength of the curve. This is a particular problem <strong>for</strong> closed loops in the xy-plane(<strong>and</strong> also <strong>for</strong> some open curves). In such cases the path may be subdivided intoshorter line segments along which one coordinate is a single-valued function ofthe other two. The sum of the line integrals along these segments is then equalto the line integral along the entire curve C. A better solution, however, is torepresent the curve in a parametric <strong>for</strong>m r(u) that is valid <strong>for</strong> its entire length.380


11.1 LINE INTEGRALS◮Evaluate the line integral I = ∮ xdy,whereC is the circle in the xy-plane defined byCx 2 + y 2 = a 2 , z =0.Adopting the usual convention mentioned above, the circle C is to be traversed in theanticlockwise direction. Taking the circle as a whole means x is not a single-valuedfunction of y. We must there<strong>for</strong>e divide the path into two parts with x =+ √ a 2 − y 2 <strong>for</strong>the semicircle lying to the right of x =0,<strong>and</strong>x = − √ a 2 − y 2 <strong>for</strong> the semicircle lying tothe left of x = 0. The required line integral is then the sum of the integrals along the twosemicircles. Substituting <strong>for</strong> x, itisgivenby∮ ∫ a √∫ −aI = xdy = a2 − y 2 dy +(− √ )a 2 − y 2 dyC−aa∫ a √=4 a2 − y 2 dy = πa 2 .0Alternatively, we can represent the entire circle parametrically, in terms of the azimuthalangle φ, sothatx = a cos φ <strong>and</strong> y = a sin φ with φ running from 0 to 2π. The integral canthere<strong>for</strong>e be evaluated over the whole circle at once. Noting that dy = a cos φdφ,wecanrewrite the line integral completely in terms of the parameter φ <strong>and</strong> obtain∮ ∫ 2πI = xdy = a 2 cos 2 φdφ= πa 2 . ◭C011.1.2 Physical examples of line integralsThere are many physical examples of line integrals, but perhaps the most commonis the expression <strong>for</strong> the total work done by a <strong>for</strong>ce F when it moves its pointof application from a point A to a point B along a given curve C. We allow themagnitude <strong>and</strong> direction of F to vary along the curve. Let the <strong>for</strong>ce act at a pointr <strong>and</strong> consider a small displacement dr along the curve; then the small amountof work done is dW = F · dr, as discussed in subsection 7.6.1 (note that dW canbe either positive or negative). There<strong>for</strong>e, the total work done in traversing thepath C is∫W C = F · dr.CNaturally, other physical quantities can be expressed in such a way. For example,the electrostatic potential energy gained by moving a charge q along a path C inan electric field E is −q ∫ CE · dr. We may also note that Ampère’s law concerningthe magnetic field B associated with a current-carrying wire can be written as∮B · dr = µ 0 I,Cwhere I is the current enclosed by a closed path C traversed in a right-h<strong>and</strong>edsense with respect to the current direction.Magnetostatics also provides a physical example of the third type of line381


LINE, SURFACE AND VOLUME INTEGRALSintegral in (11.1). If a loop of wire C carrying a current I is placed in a magneticfield B then the <strong>for</strong>ce dF on a small length dr of the wire is given by dF = Idr×B,<strong>and</strong> so the total (vector) <strong>for</strong>ce on the loop is∮F = I dr × B.C11.1.3 Line integrals with respect to a scalarIn addition to those listed in (11.1), we can <strong>for</strong>m other types of line integral,which depend on a particular curve C but <strong>for</strong> which we integrate with respectto a scalar du, rather than the vector differential dr. This distinction is somewhatarbitrary, however, since we can always rewrite line integrals containing the vectordifferential dr as a line integral with respect to some scalar parameter. If the pathC along which the integral is taken is described parametrically by r(u) thendr = drdu du,<strong>and</strong> the second type of line integral in (11.1), <strong>for</strong> example, can be written as∫C∫a · dr =Ca · drdu du.A similar procedure can be followed <strong>for</strong> the other types of line integral in (11.1).Commonly occurring special cases of line integrals with respect to a scalar are∫ ∫φds, a ds,Cwhere s is the arc length along the curve C. We can always represent C parametricallyby r(u), <strong>and</strong> from section 10.3 we have√drds =du · drdu du.The line integrals can there<strong>for</strong>e be expressed entirely in terms of the parameter u<strong>and</strong> thence evaluated.◮Evaluate the line integral I = ∫ C (x − y)2 ds, whereC is the semicircle of radius a runningfrom A =(a, 0) to B =(−a, 0) <strong>and</strong> <strong>for</strong> which y ≥ 0.The semicircular path from A to B can be described in terms of the azimuthal angle φ(measured from the x-axis) byr(φ) =a cos φ i + a sin φ j,where φ runs from 0 to π. There<strong>for</strong>e the element of arc length is given, from section 10.3,byds =√drdφ · drdφ dφ = a(cos2 φ +sin 2 φ) dφ = adφ.382C


11.2 CONNECTIVITY OF REGIONS(a) (b) (c)Figure 11.2 (a) A simply connected region; (b) a doubly connected region;(c) a triply connected region.Since (x − y) 2 = a 2 (1 − sin 2φ), the line integral becomes∫∫ πI = (x − y) 2 ds = a 3 (1 − sin 2φ) dφ = πa 3 . ◭C0As discussed in the previous chapter, the expression (10.58) <strong>for</strong> the square ofthe element of arc length in three-dimensional orthogonal curvilinear coordinatesu 1 ,u 2 ,u 3 is(ds) 2 = h 2 1 (du 1 ) 2 + h 2 2 (du 2 ) 2 + h 2 3 (du 3 ) 2 ,where h 1 ,h 2 ,h 3 are the scale factors of the coordinate system. If a curve C inthree dimensions is given parametrically by the equations u i = u i (λ) <strong>for</strong>i =1, 2, 3then the element of arc length along the curve isds =√h 2 1( ) 2 ( ) 2 ( ) 2 du1+ h 2 du22+ h 2 du33dλ.dλ dλ dλ11.2 Connectivity of regionsIn physical systems it is usual to define a scalar or vector field in some region R.In the next <strong>and</strong> some later sections we will need the concept of the connectivityof such a region in both two <strong>and</strong> three dimensions.We begin by discussing planar regions. A plane region R is said to be simplyconnected if every simple closed curve within R can be continuously shrunk toa point without leaving the region (see figure 11.2(a)). If, however, the regionR contains a hole then there exist simple closed curves that cannot be shrunkto a point without leaving R (see figure 11.2(b)). Such a region is said to bedoubly connected, since its boundary has two distinct parts. Similarly, a regionwith n − 1 holes is said to be n-fold connected, ormultiply connected (the regionin figure 11.2(c) is triply connected).383


LINE, SURFACE AND VOLUME INTEGRALSydVSRUcTCa b xFigure 11.3 A simply connected region R bounded by the curve C.These ideas can be extended to regions that are not planar, such as generalthree-dimensional surfaces <strong>and</strong> volumes. The same criteria concerning the shrinkingof closed curves to a point also apply when deciding the connectivity of suchregions. In these cases, however, the curves must lie in the surface or volumein question. For example, the interior of a torus is not simply connected, sincethere exist closed curves in the interior that cannot be shrunk to a point withoutleaving the torus. The region between two concentric spheres of different radii issimply connected.11.3 Green’s theorem in a planeIn subsection 11.1.1 we considered (amongst other things) the evaluation of lineintegrals <strong>for</strong> which the path C is closed <strong>and</strong> lies entirely in the xy-plane. Sincethe path is closed it will enclose a region R of the plane. We now discuss how toexpress the line integral around the loop as a double integral over the enclosedregion R.Suppose the functions P (x, y), Q(x, y) <strong>and</strong> their partial derivatives are singlevalued,finite <strong>and</strong> continuous inside <strong>and</strong> on the boundary C of some simplyconnected region R in the xy-plane. Green’s theorem in a plane (sometimes calledthe divergence theorem in two dimensions) then states∮∫∫ ( ∂Q(P dx+ Qdy)=CR ∂x − ∂P )dx dy, (11.4)∂y<strong>and</strong> so relates the line integral around C to a double integral over the enclosedregion R. This theorem may be proved straight<strong>for</strong>wardly in the following way.Consider the simply connected region R in figure 11.3, <strong>and</strong> let y = y 1 (x) <strong>and</strong>384


11.3 GREEN’S THEOREM IN A PLANEy = y 2 (x) be the equations of the curves STU <strong>and</strong> SVU respectively. We thenwrite∫∫∫∂Pb ∫ y2(x)R ∂y dx dy = dx dy ∂P ∫ b [ ] y=y2(x)a y 1(x) ∂y = dx P (x, y)ay=y 1(x)∫ b []= P (x, y 2 (x)) − P (x, y 1 (x)) dxa∫ b∫ a∮= − P (x, y 1 (x)) dx − P (x, y 2 (x)) dx = − Pdx.abCIf we now let x = x 1 (y) <strong>and</strong>x = x 2 (y) be the equations of the curves TSV <strong>and</strong>TUV respectively, we can similarly show that∫∫∫∂Qd ∫ x2(y)R ∂x dx dy = dy dx ∂Q ∫ d [ ] x=x2(y)c x 1(y) ∂x = dy Q(x, y)cx=x 1(y)∫ d []= Q(x 2 (y),y) − Q(x 1 (y),y) dyc∫ c∫ d∮= Q(x 1 ,y) dy + Q(x 2 ,y) dy = Qdy.dcCSubtracting these two results gives Green’s theorem in a plane.◮Show ∮ that the area of a region R enclosed by a simple closed curve C is given by A =1(xdy−ydx)=∮ xdy = − ∮ ydx. Hence calculate the area of the ellipse x = a cos φ,2 C C Cy = b sin φ.In Green’s theorem (11.4) put P = −y <strong>and</strong> Q = x; then∮∫∫∫∫(xdy− ydx)= (1+1)dx dy =2 dx dy =2A.CRR∮There<strong>for</strong>e the area of the region is A = 1 (xdy−ydx). Alternatively, we could put P =02 C<strong>and</strong> Q = x <strong>and</strong> obtain A = ∮ xdy, or put P = −y <strong>and</strong> Q = 0, which gives A = − ∮ ydx.C CThe area of the ellipse x = a cos φ, y = b sin φ is given byA = 1 ∮ydx)=2 C(xdy− 1 ∫ 2πab(cos 2 φ +sin 2 φ) dφ2 0= ab ∫ 2πdφ = πab. ◭2It may further be shown that Green’s theorem in a plane is also valid <strong>for</strong>multiply connected regions. In this case, the line integral must be taken overall the distinct boundaries of the region. Furthermore, each boundary must betraversed in the positive direction, so that a person travelling along it in thisdirection always has the region R on their left. In order to apply Green’s theorem3850


LINE, SURFACE AND VOLUME INTEGRALSyRC 2C 1Figure 11.4 A doubly connected region R bounded by the curves C 1 <strong>and</strong> C 2 .xto the region R shown in figure 11.4, the line integrals must be taken overboth boundaries, C 1 <strong>and</strong> C 2 , in the directions indicated, <strong>and</strong> the results addedtogether.We may also use Green’s theorem in a plane to investigate the path independence(or not) of line integrals when the paths lie in the xy-plane. Let us considerthe line integralI =∫ BA(P dx+ Qdy).For the line integral from A to B to be independent of the path taken, it musthave the same value along any two arbitrary paths C 1 <strong>and</strong> C 2 joining the points.Moreover, if we consider as the path the closed loop C <strong>for</strong>med by C 1 − C 2 thenthe line integral around this loop must be zero. From Green’s theorem in a plane,(11.4), we see that a sufficient condition <strong>for</strong> I = 0 is that∂P∂y = ∂Q∂x , (11.5)throughout some simply connected region R containing the loop, where we assumethat these partial derivatives are continuous in R.It may be shown that (11.5) is also a necessary condition <strong>for</strong> I = 0 <strong>and</strong> isequivalent to requiring Pdx+ Qdy to be an exact differential of some functionφ(x, y) such that Pdx+ Qdy = dφ. It follows that ∫ BA(Pdx+ Qdy)=φ(B) − φ(A)<strong>and</strong> that ∮ C(Pdx+ Qdy) around any closed loop C in the region R is identicallyzero. These results are special cases of the general results <strong>for</strong> paths in threedimensions, which are discussed in the next section.386


11.4 CONSERVATIVE FIELDS AND POTENTIALS◮Evaluate the line integral∮I = [(e x y +cosx sin y) dx +(e x +sinx cos y) dy] ,around the ellipse x 2 /a 2 + y 2 /b 2 =1.CClearly, it is not straight<strong>for</strong>ward to calculate this line integral directly. However, if we letP = e x y +cosx sin y <strong>and</strong> Q = e x +sinx cos y,then ∂P/∂y = e x +cosx cos y = ∂Q/∂x, <strong>and</strong>soPdx+ Qdy is an exact differential (itis actually the differential of the function f(x, y) =e x y +sinx sin y). From the abovediscussion, we can conclude immediately that I =0.◭11.4 Conservative fields <strong>and</strong> potentialsSo far we have made the point that, in general, the value of a line integralbetween two points A <strong>and</strong> B depends on the path C taken from A to B. Intheprevious section, however, we saw that, <strong>for</strong> paths in the xy-plane, line integralswhose integr<strong>and</strong>s have certain properties are independent of the path taken. Wenow extend that discussion to the full three-dimensional case.For line integrals of the <strong>for</strong>m ∫ Ca · dr, there exists a class of vector fields <strong>for</strong>which the line integral between two points is independent of the path taken. Suchvector fields are called conservative. A vector field a that has continuous partialderivatives in a simply connected region R is conservative if, <strong>and</strong> only if, any ofthe following is true.(i) The integral ∫ BAa · dr, whereA <strong>and</strong> B lie in the region R, is independent ofthe path from A to B. Hence the integral ∮ Ca · dr around any closed loopin R is zero.(ii) There exists a single-valued function φ of position such that a = ∇φ.(iii) ∇×a = 0.(iv) a · dr is an exact differential.The validity or otherwise of any of these statements implies the same <strong>for</strong> theother three, as we will now show.First, let us assume that (i) above is true. If the line integral from A to Bis independent of the path taken between the points then its value must be afunction only of the positions of A <strong>and</strong> B. We may there<strong>for</strong>e write∫ BAa · dr = φ(B) − φ(A), (11.6)which defines a single-valued scalar function of position φ. If the points A <strong>and</strong> Bare separated by an infinitesimal displacement dr then (11.6) becomesa · dr = dφ,387


LINE, SURFACE AND VOLUME INTEGRALSwhich shows that we require a · dr to be an exact differential: condition (iv). From(10.27) we can write dφ = ∇φ · dr, <strong>and</strong> so we have(a −∇φ) · dr =0.Since dr is arbitrary, we find that a = ∇φ; this immediately implies ∇×a = 0,condition (iii) (see (10.37)).Alternatively, if we suppose that there exists a single-valued function of positionφ such that a = ∇φ then ∇×a = 0 follows as be<strong>for</strong>e. The line integral around aclosed loop then becomes∮C∮∮a · dr = ∇φ · dr =CSince we defined φ to be single-valued, this integral is zero as required.Now suppose ∇×a = 0. From Stoke’s theorem, which is discussed in section11.9, we immediately obtain ∮ Ca · dr =0;thena = ∇φ <strong>and</strong> a · dr = dφ followas above.Finally, let us suppose a · dr = dφ. Then immediately we have a = ∇φ, <strong>and</strong>theother results follow as above.◮Evaluate the line integral I = ∫ Ba · dr, wherea A =(xy2 + z)i +(x 2 y +2)j + xk, A is thepoint (c, c, h) <strong>and</strong> B is the point (2c, c/2,h), along the different paths(i) C 1 ,givenbyx = cu, y = c/u, z = h,(ii) C 2 ,givenby2y =3c − x, z = h.Show that the vector field a is in fact conservative, <strong>and</strong> find φ such that a = ∇φ.Exp<strong>and</strong>ing out the integr<strong>and</strong>, we haveI =∫ (2c, c/2,h)(c, c, h)dφ.[(xy 2 + z) dx +(x 2 y +2)dy + xdz ] , (11.7)which we must evaluate along each of the paths C 1 <strong>and</strong> C 2 .(i) Along C 1 we have dx = cdu, dy = −(c/u 2 ) du, dz = 0, <strong>and</strong> on substituting in (11.7)<strong>and</strong> finding the limits on u, weobtain∫ 2I = c(h − 2 )du = c(h − 1).u 21(ii) Along C 2 we have 2 dy = −dx, dz = 0 <strong>and</strong>, on substituting in (11.7) <strong>and</strong> using thelimits on x, weobtainI =∫ 2cc( 12 x3 − 9 4 cx2 + 9 4 c2 x + h − 1 ) dx = c(h − 1).Hence the line integral has the same value along paths C 1 <strong>and</strong> C 2 . Taking the curl of a,we have∇×a =(0− 0)i +(1− 1)j +(2xy − 2xy)k = 0,so a is a conservative vector field, <strong>and</strong> the line integral between two points must be388


11.5 SURFACE INTEGRALSindependent of the path taken. Since a is conservative, we can write a = ∇φ. There<strong>for</strong>e, φmust satisfy∂φ∂x = xy2 + z,which implies that φ = 1 2 x2 y 2 + zx + f(y, z) <strong>for</strong> some function f. Secondly, we require∂φ∂y = x2 y + ∂f∂y = x2 y +2,which implies f =2y + g(z). Finally, since∂φ∂z = x + ∂g∂z = x,we have g =constant=k. It can be seen that we have explicitly constructed the functionφ = 1 2 x2 y 2 + zx +2y + k. ◭The quantity φ that figures so prominently in this section is called the scalarpotential function of the conservative vector field a (which satisfies ∇×a = 0), <strong>and</strong>is unique up to an arbitrary additive constant. Scalar potentials that are multivaluedfunctions of position (but in simple ways) are also of value in describingsome physical situations, the most obvious example being the scalar magneticpotential associated with a current-carrying wire. When the integral of a fieldquantity around a closed loop is considered, provided the loop does not enclosea net current, the potential is single-valued <strong>and</strong> all the above results still hold. Ifthe loop does enclose a net current, however, our analysis is no longer valid <strong>and</strong>extra care must be taken.If, instead of being conservative, a vector field b satisfies ∇ · b =0(i.e.bis solenoidal) then it is both possible <strong>and</strong> useful, <strong>for</strong> example in the theory ofelectromagnetism, to define a vector field a such that b = ∇×a. Itmaybeshownthat such a vector field a always exists. Further, if a is one such vector field thena ′ = a + ∇ψ + c, whereψ is any scalar function <strong>and</strong> c is any constant vector, alsosatisfies the above relationship, i.e. b = ∇×a ′ . This was discussed more fully insubsection 10.8.2.11.5 Surface integralsAs with line integrals, integrals over surfaces can involve vector <strong>and</strong> scalar fields<strong>and</strong>, equally, can result in either a vector or a scalar. The simplest case involvesentirely scalars <strong>and</strong> is of the <strong>for</strong>m∫φdS. (11.8)SAs analogues of the line integrals listed in (11.1), we may also encounter surfaceintegrals involving vectors, namely∫ ∫∫φdS, a · dS, a × dS. (11.9)SS389S


LINE, SURFACE AND VOLUME INTEGRALSSVdSSdSC(a)(b)Figure 11.5 (a) A closed surface <strong>and</strong> (b) an open surface. In each case anormal to the surface is shown: dS = ˆn dS.All the above integrals are taken over some surface S, which may be eitheropen or closed, <strong>and</strong> are there<strong>for</strong>e, in general, double integrals. Following thenotation <strong>for</strong> line integrals, <strong>for</strong> surface integrals over a closed surface ∫ Sis replacedby ∮ S .The vector differential dS in (11.9) represents a vector area element of thesurface S. It may also be written dS = ˆn dS, whereˆn is a unit normal to thesurface at the position of the element <strong>and</strong> dS is the scalar area of the element usedin (11.8). The convention <strong>for</strong> the direction of the normal ˆn to a surface dependson whether the surface is open or closed. A closed surface, see figure 11.5(a),does not have to be simply connected (<strong>for</strong> example, the surface of a torus is not),but it does have to enclose a volume V , which may be of infinite extent. Thedirection of ˆn is taken to point outwards from the enclosed volume as shown.An open surface, see figure 11.5(b), spans some perimeter curve C. The directionof ˆn is then given by the right-h<strong>and</strong> sense with respect to the direction in whichthe perimeter is traversed, i.e. follows the right-h<strong>and</strong> screw rule discussed insubsection 7.6.2. An open surface does not have to be simply connected but <strong>for</strong>our purposes it must be two-sided (a Möbius strip is an example of a one-sidedsurface).The <strong>for</strong>mal definition of a surface integral is very similar to that of a lineintegral. We divide the surface S into N elements of area ∆S p , p =1, 2,...,N,each with a unit normal ˆn p .If(x p ,y p ,z p ) is any point in ∆S p then the second typeof surface integral in (11.9), <strong>for</strong> example, is defined as∫∑Na · dS = lim a(x p ,y p ,z p ) · ˆn p ∆S p ,SN→∞p=1where it is required that all ∆S p → 0asN →∞.390


11.5 SURFACE INTEGRALSzkαdSSyxRdAFigure 11.6 A surface S (or part thereof) projected onto a region R in thexy-plane; dS is a surface element.11.5.1 Evaluating surface integralsWe now consider how to evaluate surface integrals over some general surface. Thisinvolves writing the scalar area element dS in terms of the coordinate differentialsof our chosen coordinate system. In some particularly simple cases this is verystraight<strong>for</strong>ward. For example, if S is the surface of a sphere of radius a (or somepart thereof) then using spherical polar coordinates θ, φ on the sphere we havedS = a 2 sin θdθdφ. For a general surface, however, it is not usually possible torepresent the surface in a simple way in any particular coordinate system. In suchcases, it is usual to work in Cartesian coordinates <strong>and</strong> consider the projections ofthe surface onto the coordinate planes.Consider a surface (or part of a surface) S as in figure 11.6. The surface S isprojected onto a region R of the xy-plane, so that an element of surface area dSprojects onto the area element dA. From the figure, we see that dA = | cos α| dS,where α is the angle between the unit vector k in the z-direction <strong>and</strong> the unitnormal ˆn to the surface at P . So, at any given point of S, we have simplydS =dA| cos α| = dA|ˆn · k| .Now, if the surface S is given by the equation f(x, y, z) = 0 then, as shown in subsection10.7.1, the unit normal at any point of the surface is given by ˆn = ∇f/|∇f|evaluated at that point, cf. (10.32). The scalar element of surface area then becomesdS =dA |∇f| dA |∇f| dA= =|ˆn · k| ∇f · k ∂f/∂z , (11.10)391


LINE, SURFACE AND VOLUME INTEGRALSwhere |∇f| <strong>and</strong> ∂f/∂z are evaluated on the surface S. We can there<strong>for</strong>e expressany surface integral over S as a double integral over the region R in the xy-plane.◮Evaluate the surface integral I = ∫ a · dS, wherea = xi <strong>and</strong> S is the surface of theShemisphere x 2 + y 2 + z 2 = a 2 with z ≥ 0.The surface of the hemisphere is shown in figure 11.7. In this case dS may be easilyexpressed in spherical polar coordinates as dS = a 2 sin θdθdφ, <strong>and</strong> the unit normal to thesurface at any point is simply ˆr. On the surface of the hemisphere we have x = a sin θ cos φ<strong>and</strong> soa · dS = x (i · ˆr) dS =(a sin θ cos φ)(sin θ cos φ)(a 2 sin θdθdφ).There<strong>for</strong>e, inserting the correct limits on θ <strong>and</strong> φ, we have∫∫ π/2I = a · dS = a 3 dθ sin 3 θS0∫ 2π0dφ cos 2 φ = 2πa33 .We could, however, follow the general prescription above <strong>and</strong> project the hemisphere Sonto the region R in the xy-plane that is a circle of radius a centred at the origin. Writingthe equation of the surface of the hemisphere as f(x, y) =x 2 + y 2 + z 2 − a 2 =0<strong>and</strong>using(11.10), we have∫ ∫∫|∇f| dAI = a · dS = x (i · ˆr) dS = x (i · ˆr)SSR ∂f/∂z .Now ∇f =2xi +2yj +2zk =2r, soonthesurfaceS we have |∇f| =2|r| =2a. OnS wealso have ∂f/∂z =2z =2 √ a 2 − x 2 − y 2 <strong>and</strong> i · ˆr = x/a. There<strong>for</strong>e, the integral becomes∫∫I =Rx 2√ dx dy.a2 − x 2 − y2 Although this integral may be evaluated directly, it is quicker to trans<strong>for</strong>m to plane polarcoordinates:∫∫ρ 2 cos 2 φI = √R ′ a2 − ρ ρdρdφ2=∫ 2π0∫ acos 2 φdφ0Making the substitution ρ = a sin u, we finally obtainI =∫ 2π0cos 2 φdφ∫ π/20ρ 3 dρ√a2 − ρ 2 .a 3 sin 3 udu= 2πa33 . ◭In the above discussion we assumed that any line parallel to the z-axis intersectsS only once. If this is not the case, we must split up the surface into smallersurfaces S 1 , S 2 etc. that are of this type. The surface integral over S is then thesum of the surface integrals over S 1 , S 2 <strong>and</strong> so on. This is always necessary <strong>for</strong>closed surfaces.Sometimes we may need to project a surface S (or some part of it) onto thezx- oryz-plane, rather than the xy-plane; <strong>for</strong> such cases, the above analysis iseasily modified.392


11.5 SURFACE INTEGRALSSzadSxaCadA = dx dyyFigure 11.7 The surface of the hemisphere x 2 + y 2 + z 2 = a 2 , z ≥ 0.11.5.2 Vector areas of surfacesThe vector area of a surface S is defined as∫S = dS,where the surface integral may be evaluated as above.S◮Find the vector area of the surface of the hemisphere x 2 + y 2 + z 2 = a 2 with z ≥ 0.As in the previous example, dS = a 2 sin θdθdφˆr in spherical polar coordinates. There<strong>for</strong>ethe vector area is given by∫∫S = a 2 sin θ ˆr dθ dφ.SNow, since ˆr varies over the surface S, it also must be integrated. This is most easilyachieved by writing ˆr in terms of the constant Cartesian basis vectors. On S we haveˆr =sinθ cos φ i +sinθ sin φ j +cosθ k,so the expression <strong>for</strong> the vector area becomes( ∫ 2π ∫ ) (π/2∫ 2π ∫ )π/2S = i a 2 cos φdφ sin 2 θdθ + j a 2 sin φdφ sin 2 θdθ0000( ∫ 2π ∫ )π/2+ k a 2 dφ sin θ cos θdθ= 0 + 0 + πa 2 k = πa 2 k.00Note that the magnitude of S is the projected area, of the hemisphere onto the xy-plane,<strong>and</strong> not the surface area of the hemisphere. ◭393


LINE, SURFACE AND VOLUME INTEGRALSdrCrOFigure 11.8 The conical surface spanning the perimeter C <strong>and</strong> having itsvertex at the origin.The hemispherical shell discussed above is an example of an open surface. Fora closed surface, however, the vector area is always zero. This may be seen byprojecting the surface down onto each Cartesian coordinate plane in turn. Foreach projection, every positive element of area on the upper surface is cancelledby the corresponding negative element on the lower surface. There<strong>for</strong>e, eachcomponent of S = ∮ SdS vanishes.An important corollary of this result is that the vector area of an open surfacedepends only on its perimeter, or boundary curve, C. This may be proved asfollows. If surfaces S 1 <strong>and</strong> S 2 have the same perimeter then S 1 − S 2 is a closedsurface, <strong>for</strong> which∮ ∫ ∫dS = dS − dS = 0.S 1 S 2Hence S 1 = S 2 . Moreover, we may derive an expression <strong>for</strong> the vector area ofan open surface S solely in terms of a line integral around its perimeter C.Since we may choose any surface with perimeter C, we will consider a conewith its vertex at the origin (see figure 11.8). The vector area of the elementarytriangular region shown in the figure is dS = 1 2r × dr. There<strong>for</strong>e, the vector areaof the cone, <strong>and</strong> hence of any open surface with perimeter C, is given by the lineintegralS = 1 ∮r × dr.2CFor a surface confined to the xy-plane, r = xi + yj <strong>and</strong> dr = dx i + dy j,<strong>and</strong> we∮obtain <strong>for</strong> this special case that the area of the surface is given byA = 1 2 C(xdy− ydx), as we found in section 11.3.394


11.5 SURFACE INTEGRALS◮Find the vector area of the surface ∮ of the hemisphere x 2 + y 2 + z 2evaluating the line integral S = 1 r × dr around its perimeter.2 C= a 2 , z ≥ 0, byThe perimeter C of the hemisphere is the circle x 2 + y 2 = a 2 , on which we haver = a cos φ i + a sin φ j, dr = −a sin φdφi + a cos φdφj.There<strong>for</strong>e the cross product r × dr is given byi j kr × dr =a cos φ asin φ 0∣ −a sin φdφ acos φdφ 0 ∣ = a2 (cos 2 φ +sin 2 φ) dφ k = a 2 dφ k,<strong>and</strong> the vector area becomes∫ 2πS = 1 2 a2 k dφ = πa 2 k. ◭011.5.3 Physical examples of surface integralsThere are many examples of surface integrals in the physical sciences. Surfaceintegrals of the <strong>for</strong>m (11.8) occur in computing the total electric charge on asurface or the mass of a shell, ∫ Sρ(r) dS, given the charge or mass density ρ(r).For surface integrals involving vectors, the second <strong>for</strong>m in (11.9) is the mostcommon. For a vector field a, the surface integral ∫ Sa · dS is called the fluxof a through S. Examples of physically important flux integrals are numerous.For example, let us consider a surface S in a fluid with density ρ(r) that has avelocity field v(r). The mass of fluid crossing an element of surface area dS intime dt is dM = ρv · dS dt. There<strong>for</strong>e the net total mass flux of fluid crossing Sis M = ∫ Sρ(r)v(r) · dS. As a another example, the electromagnetic flux of energyout of a given volume V bounded by a surface S is ∮ S(E × H) · dS.The solid angle, to be defined below, subtended at a point O by a surface (closedor otherwise) can also be represented by an integral of this <strong>for</strong>m, although it isnot strictly a flux integral (unless we imagine isotropic rays radiating from O).The integral∫ ∫r · dS ˆr · dSΩ=S r 3 =S r 2 , (11.11)gives the solid angle Ω subtended at O by a surface S if r is the position vectormeasured from O of an element of the surface. A little thought will show that(11.11) takes account of all three relevant factors: the size of the element ofsurface, its inclination to the line joining the element to O <strong>and</strong> the distance fromO. Such a general expression is often useful <strong>for</strong> computing solid angles when thethree-dimensional geometry is complicated. Note that (11.11) remains valid whenthe surface S is not convex <strong>and</strong> when a single ray from O in certain directionswould cut S in more than one place (but we exclude multiply connected regions).395


LINE, SURFACE AND VOLUME INTEGRALSIn particular, when the surface is closed Ω = 0 if O is outside S <strong>and</strong> Ω = 4π if Ois an interior point.Surface integrals resulting in vectors occur less frequently. An example isaf<strong>for</strong>ded, however, by the total resultant <strong>for</strong>ce experienced by a body immersed ina stationary fluid in which the hydrostatic pressure is given by p(r). The pressureis everywhere inwardly directed <strong>and</strong> the resultant <strong>for</strong>ce is F = − ∮ SpdS, takenover the whole surface.11.6 Volume integralsVolume integrals are defined in an obvious way <strong>and</strong> are generally simpler thanline or surface integrals since the element of volume dV is a scalar quantity. Wemay encounter volume integrals of the <strong>for</strong>ms∫∫φdV, a dV . (11.12)VClearly, the first <strong>for</strong>m results in a scalar, whereas the second <strong>for</strong>m yields a vector.Two closely related physical examples, one of each kind, are provided by the totalmass of a fluid contained in a volume V , given by ∫ Vρ(r) dV , <strong>and</strong> the total linearmomentum of that same fluid, given by ∫ Vρ(r)v(r) dV where v(r) is the velocityfield in the fluid. As a slightly more complicated example of a volume integral wemay consider the following.◮Find an expression <strong>for</strong> the angular momentum of a solid body rotating with angularvelocity ω about an axis through the origin.Consider a small volume element dV situated at position r; its linear momentum is ρdVṙ,where ρ = ρ(r) is the density distribution, <strong>and</strong> its angular momentum about O is r × ρṙ dV .Thus <strong>for</strong> the whole body the angular momentum L is∫L = (r × ṙ)ρdV.Putting ṙ = ω × r yields∫∫L = [r × (ω × r)] ρdV =VVVV∫ωr 2 ρdV − (r · ω)rρdV. ◭VThe evaluation of the first type of volume integral in (11.12) has already beenconsidered in our discussion of multiple integrals in chapter 6. The evaluation ofthe second type of volume integral follows directly since we can write∫ ∫ ∫∫a dV = i a x dV + j a y dV + k a z dV , (11.13)VVwhere a x ,a y ,a z are the Cartesian components of a. Of course, we could havewritten a in terms of the basis vectors of some other coordinate system (e.g.spherical polars) but, since such basis vectors are not, in general, constant, they396VV


11.6 VOLUME INTEGRALSdSSVrOFigure 11.9 A general volume V containing the origin <strong>and</strong> bounded by theclosed surface S.cannot be taken out of the integral sign as in (11.13) <strong>and</strong> must be included aspart of the integr<strong>and</strong>.11.6.1 Volumes of three-dimensional regionsAs discussed in chapter 6, the volume of a three-dimensional region V is simplyV = ∫ VdV , which may be evaluated directly once the limits of integration havebeen found. However, the volume of the region obviously depends only on thesurface S that bounds it. We should there<strong>for</strong>e be able to express the volume Vin terms of a surface integral over S. This is indeed possible, <strong>and</strong> the appropriateexpression may derived as follows. Referring to figure 11.9, let us suppose thatthe origin O is contained within V . The volume of the small shaded cone isdV = 1 3r · dS; the total volume of the region is thus given byV = 1 ∮r · dS.3 SIt may be shown that this expression is valid even when O is not contained in V .Although this surface integral <strong>for</strong>m is available, in practice it is usually simplerto evaluate the volume integral directly.◮Find the volume enclosed between a sphere of radius a centred on the origin <strong>and</strong> a circularcone of half-angle α with its vertex at the origin.The element of vector area dS on the surface of the sphere is given in spherical polarcoordinates by a 2 sin θdθdφˆr. Now taking the axis of the cone to lie along the z-axis (fromwhich θ is measured) the required volume is given byV = 1 ∮r · dS = 1 ∫ 2π ∫ αdφ a 2 sin θ r · ˆr dθ33S= 1 30∫ 2π0dφ0∫ α0a 3 sin θdθ= 2πa3 (1 − cos α). ◭3397


LINE, SURFACE AND VOLUME INTEGRALS11.7 Integral <strong>for</strong>ms <strong>for</strong> grad, div <strong>and</strong> curlIn the previous chapter we defined the vector operators grad, div <strong>and</strong> curl in purelymathematical terms, which depended on the coordinate system in which they wereexpressed. An interesting application of line, surface <strong>and</strong> volume integrals is theexpression of grad, div <strong>and</strong> curl in coordinate-free, geometrical terms. If φ is ascalar field <strong>and</strong> a is a vector field then it may be shown that at any point P( ∮ 1∇φ = limV →0 V S( ∮ 1∇ · a = limV →0 V∇×a = limV →0( 1V)φdS)a · dSS∮ )dS × aS(11.14)(11.15)(11.16)where V is a small volume enclosing P <strong>and</strong> S is its bounding surface. Indeed,we may consider these equations as the (geometrical) definitions of grad, div <strong>and</strong>curl. An alternative, but equivalent, geometrical definition of ∇×a at a point P ,which is often easier to use than (11.16), is given by( ∮ )1(∇×a) · ˆn = lim a · dr , (11.17)A→0 A Cwhere C is a plane contour of area A enclosing the point P <strong>and</strong> ˆn is the unitnormal to the enclosed planar area.It may be shown, in any coordinate system, that all the above equations areconsistent with our definitions in the previous chapter, although the difficulty ofproof depends on the chosen coordinate system. The most general coordinatesystem encountered in that chapter was one with orthogonal curvilinear coordinatesu 1 ,u 2 ,u 3 , of which Cartesians, cylindrical polars <strong>and</strong> spherical polars are allspecial cases. Although it may be shown that (11.14) leads to the usual expression<strong>for</strong> grad in curvilinear coordinates, the proof requires complicated manipulationsof the derivatives of the basis vectors with respect to the coordinates <strong>and</strong> is notpresented here. In Cartesian coordinates, however, the proof is quite simple.◮Show that the geometrical definition of grad leads to the usual expression <strong>for</strong> ∇φ inCartesian coordinates.Consider the surface S of a small rectangular volume element ∆V =∆x ∆y ∆z that has itsfaces parallel to the x, y, <strong>and</strong>z coordinate surfaces; the point P (see above) is at one corner.We must calculate the surface integral (11.14) over each of its six faces. Remembering thatthe normal to the surface points outwards from the volume on each face, the two faceswith x = constant have areas ∆S = −i ∆y ∆z <strong>and</strong> ∆S = i ∆y ∆z respectively. Furthermore,over each small surface element, we may take φ to be constant, so that the net contribution398


11.7 INTEGRAL FORMS FOR grad, div AND curlto the surface integral from these two faces is then([(φ +∆φ) − φ]∆y ∆z i = φ + ∂φ )∂x ∆x − φ ∆y ∆z i= ∂φ ∆x ∆y ∆z i.∂xThe surface integral over the pairs of faces with y =constant<strong>and</strong>z = constant respectivelymay be found in a similar way, <strong>and</strong> we obtain∮ ( ∂φφdS =S ∂x i + ∂φ∂y j + ∂φ )∂z k ∆x ∆y ∆z.There<strong>for</strong>e ∇φ at the point P is given by[ (1 ∂φ∇φ = lim∆x,∆y,∆z→0 ∆x ∆y ∆z ∂x i + ∂φ∂y j + ∂φ ) ]∂z k ∆x ∆y ∆z= ∂φ∂x i + ∂φ∂y j + ∂φ∂z k. ◭We now turn to (11.15) <strong>and</strong> (11.17). These geometrical definitions may beshown straight<strong>for</strong>wardly to lead to the usual expressions <strong>for</strong> div <strong>and</strong> curl inorthogonal curvilinear coordinates.◮By considering the infinitesimal volume element dV = h 1 h 2 h 3 ∆u 1 ∆u 2 ∆u 3 shown in figure11.10, show that (11.15) leads to the usual expression <strong>for</strong> ∇·a in orthogonal curvilinearcoordinates.Let us write the vector field in terms of its components with respect to the basis vectorsof the curvilinear coordinate system as a = a 1 ê 1 + a 2 ê 2 + a 3 ê 3 . We consider first thecontribution to the RHS of (11.15) from the two faces with u 1 = constant, i.e. PQRS<strong>and</strong> the face opposite it (see figure 11.10). Now, the volume element is <strong>for</strong>med from theorthogonal vectors h 1 ∆u 1 ê 1 , h 2 ∆u 2 ê 2 <strong>and</strong> h 3 ∆u 3 ê 3 at the point P <strong>and</strong>so<strong>for</strong>PQRS wehave∆S = h 2 h 3 ∆u 2 ∆u 3 ê 3 × ê 2 = −h 2 h 3 ∆u 2 ∆u 3 ê 1 .Reasoning along the same lines as in the previous example, we conclude that the contributionto the surface integral of a · dS over PQRS <strong>and</strong> its opposite face taken together isgiven by∂(a · ∆S)∆u 1 =∂ (a 1 h 2 h 3 )∆u 1 ∆u 2 ∆u 3 .∂u 1 ∂u 1The surface integrals over the pairs of faces with u 2 = constant <strong>and</strong> u 3 = constantrespectively may be found in a similar way, <strong>and</strong> we obtain∮ [ ∂a · dS = (a 1 h 2 h 3 )+∂ (a 2 h 3 h 1 )+∂u 1 ∂u 2SThere<strong>for</strong>e ∇ · a at the point P is given by[1∇ · a = lim∆u 1 ,∆u 2 ,∆u 3 →0 h 1 h 2 h 3 ∆u 1 ∆u 2 ∆u 3∮S= 1 [ ∂(a 1 h 2 h 3 )+h 1 h 2 h 3 ∂u 1∂∂u 3(a 3 h 1 h 2 )∂∂u 2(a 2 h 3 h 1 )+]a · dS]∆u 1 ∆u 2 ∆u 3 .∂ ](a 3 h 1 h 2 ) . ◭∂u 3399


LINE, SURFACE AND VOLUME INTEGRALSzh 1 ∆u 1 ê 1RS TQh 2 ∆u 2 ê 2P h 3 ∆u 3 ê 3yxFigure 11.10 A general volume ∆V in orthogonal curvilinear coordinatesu 1 ,u 2 ,u 3 . PT gives the vector h 1 ∆u 1 ê 1 , PS gives h 2 ∆u 2 ê 2 <strong>and</strong> PQ givesh 3 ∆u 3 ê 3 .◮By considering the infinitesimal planar surface element PQRS in figure 11.10, show that(11.17) leads to the usual expression <strong>for</strong> ∇×a in orthogonal curvilinear coordinates.The planar surface PQRS is defined by the orthogonal vectors h 2 ∆u 2 ê 2 <strong>and</strong> h 3 ∆u 3 ê 3at the point P . If we traverse the loop in the direction PSRQ then, by the right-h<strong>and</strong>convention, the unit normal to the plane is ê 1 .Writinga = a 1 ê 1 + a 2 ê 2 + a 3 ê 3 , the lineintegral around the loop in this direction is given by∮PSRQa · dr = a 2 h 2 ∆u 2 +[− a 2 h 2 +[ ∂= (a 3 h 3 ) −∂u 2[a 3 h 3 +∂ ](a 3 h 3 )∆u 2 ∆u 3∂u 2∂ ](a 2 h 2 )∆u 3 ∆u 2 − a 3 h 3 ∆u 3∂u 3]∂∂u 3(a 2 h 2 )∆u 2 ∆u 3 .There<strong>for</strong>e from (11.17) the component of ∇×a in the direction ê 1 at P is given by[]1(∇×a) 1 = lima · dr∆u 2 ,∆u 3 →0 h 2 h 3 ∆u 2 ∆u 3∮PSRQ= 1 [ ∂(h 3 a 3 ) −∂ ](h 2 a 2 ) .h 2 h 3 ∂u 2 ∂u 3The other two components are found by cyclically permuting the subscripts 1, 2, 3. ◭Finally, we note that we can also write the ∇ 2 operator as a surface integral bysetting a = ∇φ in (11.15), to obtain( ∮ )1∇ 2 φ = ∇ · ∇φ = lim ∇φ · dS .V →0 V400S


11.8 DIVERGENCE THEOREM AND RELATED THEOREMS11.8 Divergence theorem <strong>and</strong> related theoremsThe divergence theorem relates the total flux of a vector field out of a closedsurface S to the integral of the divergence of the vector field over the enclosedvolume V ; it follows almost immediately from our geometrical definition ofdivergence (11.15).Imagine a volume V , in which a vector field a is continuous <strong>and</strong> differentiable,to be divided up into a large number of small volumes V i . Using (11.15), we have<strong>for</strong> each small volume∮(∇ · a)V i ≈ a · dS,S iwhere S i is the surface of the small volume V i . Summing over i we find thatcontributions from surface elements interior to S cancel since each surface elementappears in two terms with opposite signs, the outward normals in the two termsbeing equal <strong>and</strong> opposite. Only contributions from surface elements that are alsoparts of S survive. If each V i is allowed to tend to zero then we obtain thedivergence theorem,∫∮∇ · a dV = a · dS. (11.18)VSWe note that the divergence theorem holds <strong>for</strong> both simply <strong>and</strong> multiply connectedsurfaces, provided that they are closed <strong>and</strong> enclose some non-zero volumeV . The divergence theorem may also be extended to tensor fields (see chapter 26).The theorem finds most use as a tool in <strong>for</strong>mal manipulations, but sometimesit is of value in trans<strong>for</strong>ming surface integrals of the <strong>for</strong>m ∫ Sa · dS into volumeintegrals or vice versa. For example, setting a = r we immediately obtain∫∫∮∇ · r dV = 3 dV =3V = r · dS,VVwhich gives the expression <strong>for</strong> the volume of a region found in subsection 11.6.1.The use of the divergence theorem is further illustrated in the following example.◮Evaluate the surface integral I = ∫ S a · dS, wherea =(y − x) i + x2 z j +(z + x 2 ) k <strong>and</strong> Sis the open surface of the hemisphere x 2 + y 2 + z 2 = a 2 , z ≥ 0.We could evaluate this surface integral directly, but the algebra is somewhat lengthy. Wewill there<strong>for</strong>e evaluate it by use of the divergence theorem. Since the latter only holds<strong>for</strong> closed surfaces enclosing a non-zero volume V , let us first consider the closed surfaceS ′ = S + S 1 ,whereS 1 is the circular area in the xy-plane given by x 2 + y 2 ≤ a 2 , z =0;S ′then encloses a hemispherical volume V . By the divergence theorem we have∫∮ ∫ ∫∇ · a dV = a · dS = a · dS + a · dS.VS ′ SS 1Now ∇ · a = −1+0+1=0,sowecanwrite∫∫a · dS = − a · dS.SS 1401S


LINE, SURFACE AND VOLUME INTEGRALSyCRdxdrdyˆn dsxFigure 11.11 A closed curve C in the xy-plane bounding a region R. Vectorstangent <strong>and</strong> normal to the curve at a given point are also shown.The surface integral over S 1 is easily evaluated. Remembering that the normal to thesurface points outward from the volume, a surface element on S 1 is simply dS = −k dx dy.On S 1 we also have a =(y − x) i + x 2 k,sothat∫ ∫∫I = − a · dS = x 2 dx dy,S 1 Rwhere R is the circular region in the xy-plane given by x 2 + y 2 ≤ a 2 . Trans<strong>for</strong>ming to planepolar coordinates we have∫∫∫ 2π ∫ aI = ρ 2 cos 2 φ ρ dρ dφ = cos 2 φdφ ρ 3 dρ = πa4R ′ 00 4 . ◭It is also interesting to consider the two-dimensional version of the divergencetheorem. As an example, let us consider a two-dimensional planar region R inthe xy-plane bounded by some closed curve C (see figure 11.11). At any pointon the curve the vector dr = dx i + dy j is a tangent to the curve <strong>and</strong> the vectorˆn ds = dy i − dx j is a normal pointing out of the region R. If the vector field a iscontinuous <strong>and</strong> differentiable in R then the two-dimensional divergence theoremin Cartesian coordinates gives∫∫R( ∂ax∂x + ∂a y∂y)∮dx dy =∮a · ˆn ds = (a x dy − a y dx).CLetting P = −a y <strong>and</strong> Q = a x , we recover Green’s theorem in a plane, which wasdiscussed in section 11.3.11.8.1 Green’s theoremsConsider two scalar functions φ <strong>and</strong> ψ that are continuous <strong>and</strong> differentiable insome volume V bounded by a surface S. Applying the divergence theorem to the402


11.8 DIVERGENCE THEOREM AND RELATED THEOREMSvector field φ∇ψ we obtain∮∫φ∇ψ · dS =S∫=VV∇ · (φ∇ψ) dV[φ∇ 2 ψ +(∇φ) · (∇ψ) ] dV . (11.19)Reversing the roles of φ <strong>and</strong> ψ in (11.19) <strong>and</strong> subtracting the two equations gives∮∫(φ∇ψ − ψ∇φ) · dS = (φ∇ 2 ψ − ψ∇ 2 φ) dV . (11.20)SEquation (11.19) is usually known as Green’s first theorem <strong>and</strong> (11.20) as hissecond. Green’s second theorem is useful in the development of the Green’sfunctions used in the solution of partial differential equations (see chapter 21).V11.8.2 Other related integral theoremsThere exist two other integral theorems which are closely related to the divergencetheorem <strong>and</strong> which are of some use in physical applications. If φ is a scalar field<strong>and</strong> b is a vector field <strong>and</strong> both φ <strong>and</strong> b satisfy our usual differentiabilityconditions in some volume V bounded by a closed surface S then∫ ∮∇φdV = φdS, (11.21)VS∫∮∇×b dV = dS × b. (11.22)VS◮Use the divergence theorem to prove (11.21).In the divergence theorem (11.18) let a = φc, wherec is a constant vector. We then have∫∮∇ · (φc) dV = φc · dS.Exp<strong>and</strong>ing out the integr<strong>and</strong> on the LHS we have∇ · (φc) =φ∇ · c + c · ∇φ = c · ∇φ,since c is constant. Also, φc · dS = c · φdS, soweobtain∫∮c · (∇φ) dV = c · φdS.Since c is constant we may take it out of both integrals to give∫∮c · ∇φdV = c · φdS,<strong>and</strong> since c is arbitrary we obtain the stated result (11.21). ◭VVVEquation (11.22) may be proved in a similar way by letting a = b × c in thedivergence theorem, where c is again a constant vector.403SSS


LINE, SURFACE AND VOLUME INTEGRALS11.8.3 Physical applications of the divergence theoremThe divergence theorem is useful in deriving many of the most important partialdifferential equations in physics (see chapter 20). The basic idea is to use thedivergence theorem to convert an integral <strong>for</strong>m, often derived from observation,into an equivalent differential <strong>for</strong>m (used in theoretical statements).◮For a compressible fluid with time-varying position-dependent density ρ(r,t) <strong>and</strong> velocityfield v(r,t), in which fluid is neither being created nor destroyed, show that∂ρ+ ∇ · (ρv) =0.∂tFor an arbitrary volume V in the fluid, the conservation of mass tells us that the rate ofincrease or decrease of the mass M of fluid in the volume must equal the net rate at whichfluid is entering or leaving the volume, i.e.∮dM= − ρv · dS,dt Swhere S is the surface bounding V . But the mass of fluid in V is simply M = ∫ ρdV,soVwe have∫ ∮dρdV + ρv · dS =0.dt VSTaking the derivative inside the first integral on the RHS <strong>and</strong> using the divergence theoremto rewrite the second integral, we obtain∫∂ρV ∂t∫VdV + ∇ · (ρv) dV =∫V[ ∂ρ∂t + ∇ · (ρv) ]dV =0.Since the volume V is arbitrary, the integr<strong>and</strong> (which is assumed continuous) must beidentically zero, so we obtain∂ρ+ ∇ · (ρv) =0.∂tThis is known as the continuity equation. It can also be applied to other systems, <strong>for</strong>example those in which ρ is the density of electric charge or the heat content, etc. For theflow of an incompressible fluid, ρ = constant <strong>and</strong> the continuity equation becomes simply∇ · v =0.◭In the previous example, we assumed that there were no sources or sinks inthe volume V , i.e. that there was no part of V in which fluid was being createdor destroyed. We now consider the case where a finite number of point sources<strong>and</strong>/or sinks are present in an incompressible fluid. Let us first consider thesimple case where a single source is located at the origin, out of which a quantityof fluid flows radially at a rate Q (m 3 s −1 ). The velocity field is given byv =Qr4πr 3 = Qˆr4πr 2 .Now, <strong>for</strong> a sphere S 1 of radius r centred on the source, the flux across S 1 is∮v · dS = |v|4πr 2 = Q.S 1404


11.8 DIVERGENCE THEOREM AND RELATED THEOREMSSince v has a singularity at the origin it is not differentiable there, i.e. ∇·v is notdefined there, but at all other points ∇ · v = 0, as required <strong>for</strong> an incompressiblefluid. There<strong>for</strong>e, from the divergence theorem, <strong>for</strong> any closed surface S 2 that doesnot enclose the origin we have∮ ∫v · dS = ∇ · v dV =0.S 2 VThus we see that the surface integral ∮ Sv · dS has value Q or zero dependingon whether or not S encloses the source. In order that the divergence theorem isvalid <strong>for</strong> all surfaces S, irrespective of whether they enclose the source, we write∇ · v = Qδ(r),where δ(r) is the three-dimensional Dirac delta function. The properties of thisfunction are discussed fully in chapter 13, but <strong>for</strong> the moment we note that it isdefined in such a way that∫Vδ(r − a) =0 <strong>for</strong>r ≠ a,{f(a) if a lies in Vf(r)δ(r − a) dV =0 otherwise<strong>for</strong> any well-behaved function f(r). There<strong>for</strong>e, <strong>for</strong> any volume V containing thesource at the origin, we have∫∫∇ · v dV = Q δ(r) dV = Q,Vwhich is consistent with ∮ Sv · dS = Q <strong>for</strong> a closed surface enclosing the source.Hence, by introducing the Dirac delta function the divergence theorem can bemade valid even <strong>for</strong> non-differentiable point sources.The generalisation to several sources <strong>and</strong> sinks is straight<strong>for</strong>ward. For example,if a source is located at r = a <strong>and</strong> a sink at r = b then the velocity field isv =<strong>and</strong> its divergence is given byV(r − a)Q (r − b)Q−4π|r − a|34π|r − b| 3∇ · v = Qδ(r − a) − Qδ(r − b).There<strong>for</strong>e, the integral ∮ Sv · dS has the value Q if S encloses the source, −Q ifS encloses the sink <strong>and</strong> 0 if S encloses neither the source nor sink or enclosesthem both. This analysis also applies to other physical systems – <strong>for</strong> example, inelectrostatics we can regard the sources <strong>and</strong> sinks as positive <strong>and</strong> negative pointcharges respectively <strong>and</strong> replace v by the electric field E.405


LINE, SURFACE AND VOLUME INTEGRALS11.9 Stokes’ theorem <strong>and</strong> related theoremsStokes’ theorem is the ‘curl analogue’ of the divergence theorem <strong>and</strong> relates theintegral of the curl of a vector field over an open surface S to the line integral ofthe vector field around the perimeter C bounding the surface.Following the same lines as <strong>for</strong> the derivation of the divergence theorem, wecan divide the surface S into many small areas S i with boundaries C i <strong>and</strong> unitnormals ˆn i . Using (11.17), we have <strong>for</strong> each small area∮(∇×a) · ˆn i S i ≈ a · dr.C iSumming over i we find that on the RHS all parts of all interior boundariesthat are not part of C are included twice, being traversed in opposite directionson each occasion <strong>and</strong> thus contributing nothing. Only contributions from lineelements that are also parts of C survive. If each S i is allowed to tend to zerothen we obtain Stokes’ theorem,∫∮(∇×a) · dS = a · dr. (11.23)SCWe note that Stokes’ theorem holds <strong>for</strong> both simply <strong>and</strong> multiply connected opensurfaces, provided that they are two-sided. Stokes’ theorem may also be extendedto tensor fields (see chapter 26).Just as the divergence theorem (11.18) can be used to relate volume <strong>and</strong> surfaceintegrals <strong>for</strong> certain types of integr<strong>and</strong>, Stokes’ theorem can be used in evaluatingsurface integrals of the <strong>for</strong>m ∮ S(∇×a) · dS as line integrals or vice versa.◮Given the vector field a = y i − x j + z k, verify Stokes’ theorem <strong>for</strong> the hemisphericalsurface x 2 + y 2 + z 2 = a 2 , z ≥ 0.Let us first evaluate the surface integral∫S(∇×a) · dSover the hemisphere. It is easily shown that ∇×a = −2 k, <strong>and</strong> the surface element isdS = a 2 sin θdθdφ ˆr in spherical polar coordinates. There<strong>for</strong>e∫∫ 2π ∫ π/2(∇×a) · dS = dφ dθ ( −2a 2 sin θ ) ˆr · kS0∫ 2π= −2a 2 dφ∫ 2π= −2a 2 dφ000∫ π/20∫ π/20( z)sin θ dθasin θ cos θdθ= −2πa 2 .We now evaluate the line integral around the perimeter curve C of the surface, which406


11.9 STOKES’ THEOREM AND RELATED THEOREMSis the circle x 2 + y 2 = a 2 in the xy-plane. This is given by∮ ∮a · dr = (y i − x j + z k) · (dx i + dy j + dz k)CC∮= (ydx− xdy).CUsing plane polar coordinates, on C we have x = a cos φ, y = a sin φ so that dx =−a sin φdφ, dy = a cos φdφ, <strong>and</strong> the line integral becomes∮∫ 2π∫ 2π(ydx− xdy)=−a 2 (sin 2 φ +cos 2 φ) dφ = −a 2 dφ = −2πa 2 .C0Since the surface <strong>and</strong> line integrals have the same value, we have verified Stokes’ theoremin this case. ◭The two-dimensional version of Stokes’ theorem also yields Green’s theorem ina plane. Consider the region R in the xy-plane shown in figure 11.11, in which avector field a is defined. Since a = a x i+a y j, we have ∇×a =(∂a y /∂x−∂a x /∂y) k,<strong>and</strong> Stokes’ theorem becomes∫∫R( ∂ay∂x − ∂a x∂y)∮dx dy = (a x dx + a y dy).CLetting P = a x <strong>and</strong> Q = a y we recover Green’s theorem in a plane, (11.4).011.9.1 Related integral theoremsAs <strong>for</strong> the divergence theorem, there exist two other integral theorems that areclosely related to Stokes’ theorem. If φ is a scalar field <strong>and</strong> b is a vector field,<strong>and</strong> both φ <strong>and</strong> b satisfy our usual differentiability conditions on some two-sidedopen surface S bounded by a closed perimeter curve C, then∫∮dS ×∇φ = φdr, (11.24)SC∫∮(dS ×∇) × b = dr × b. (11.25)SC◮Use Stokes’ theorem to prove (11.24).In Stokes’ theorem, (11.23), let a = φc, wherec is a constant vector. We then have∫∮[∇×(φc)] · dS = φc · dr. (11.26)SExp<strong>and</strong>ing out the integr<strong>and</strong> on the LHS we have∇×(φc) =∇φ × c + φ∇×c = ∇φ × c,since c is constant, <strong>and</strong> the scalar triple product on the LHS of (11.26) can there<strong>for</strong>e bewritten[∇×(φc)] · dS =(∇φ × c) · dS = c · (dS ×∇φ).407C


LINE, SURFACE AND VOLUME INTEGRALSSubstituting this into (11.26) <strong>and</strong> taking c out of both integrals because it is constant, wefind∫∮c · dS ×∇φ = c · φdr.SCSince c is an arbitrary constant vector we there<strong>for</strong>e obtain the stated result (11.24). ◭Equation (11.25) may be proved in a similar way, by letting a = b × c in Stokes’theorem, where c is again a constant vector. We also note that by setting b = rin (11.25) we find∫∮(dS ×∇) × r = dr × r.SCExp<strong>and</strong>ing out the integr<strong>and</strong> on the LHS gives(dS ×∇) × r = dS − dS(∇ · r) =dS − 3 dS = −2 dS.There<strong>for</strong>e, as we found in subsection 11.5.2, the vector area of an open surface Sis given by∫S = dS = 1 ∮r × dr.S 2 C11.9.2 Physical applications of Stokes’ theoremLike the divergence theorem, Stokes’ theorem is useful in converting integralequations into differential equations.◮From Ampère’s law, derive Maxwell’s equation in the case where the currents are steady,i.e. ∇×B − µ 0 J = 0.Ampère’s rule <strong>for</strong> a distributed current with current density J is∮∫B · dr = µ 0 J · dS,C<strong>for</strong> any circuit C bounding a surface S. Using Stokes’ theorem, the LHS can be trans<strong>for</strong>medinto ∫ (∇×B) · dS; henceS∫S(∇×B − µ 0 J) · dS =0<strong>for</strong> any surface S. This can only be so if ∇×B − µ 0 J = 0, which is the required relation.Similarly, from Faraday’s law of electromagnetic induction we can derive Maxwell’sequation ∇×E = −∂B/∂t. ◭In subsection 11.8.3 we discussed the flow of an incompressible fluid in thepresence of several sources <strong>and</strong> sinks. Let us now consider vortex flow in anincompressible fluid with a velocity fieldSv = 1 ρêφ,in cylindrical polar coordinates ρ, φ, z. For this velocity field ∇×v equals zero408


11.10 EXERCISESeverywhere except on the axis ρ =0,wherev has a singularity. There<strong>for</strong>e ∮ C v · drequals zero <strong>for</strong> any path C that does not enclose the vortex line on the axis <strong>and</strong>2π if C does enclose the axis. In order <strong>for</strong> Stokes’ theorem to be valid <strong>for</strong> allpaths C, we there<strong>for</strong>e set∇×v =2πδ(ρ),where δ(ρ) is the Dirac delta function, to be discussed in subsection 13.1.3. Now,since ∇×v = 0, except on the axis ρ = 0, there exists a scalar potential ψ suchthat v = ∇ψ. It may easily be shown that ψ = φ, the polar angle. There<strong>for</strong>e, if Cdoes not enclose the axis then ∮ ∮v · dr = dφ =0,C<strong>and</strong> if C does enclose the axis,∮v · dr =∆φ =2πn,Cwhere n is the number of times we traverse C. Thus φ is a multivalued potential.Similar analyses are valid <strong>for</strong> other physical systems – <strong>for</strong> example, in magnetostaticswe may replace the vortex lines by current-carrying wires <strong>and</strong> the velocityfield v by the magnetic field B.11.10 Exercises11.1 The vector field F is defined byF =2xzi +2yz 2 j +(x 2 +2y 2 z − 1)k.Calculate ∇×F <strong>and</strong> deduce that F can be written F = ∇φ. Determine the <strong>for</strong>mof φ.11.2 The vector field Q is defined byQ = [ 3x 2 (y + z)+y 3 + z 3] i + [ 3y 2 (z + x)+z 3 + x 3] j + [ 3z 2 (x + y)+x 3 + y 3] k.Show that Q is a conservative field, construct its potential function <strong>and</strong> henceevaluate the integral J = ∫ Q · dr along any line connecting the point A at(1, −1, 1) to B at (2, 1, 2).11.3 F is a vector field xy 2 i+2j+xk, <strong>and</strong>L is a path parameterised by x = ct, y = c/t,z = d <strong>for</strong> the range 1 ≤ t ≤ 2. Evaluate (a) ∫ F dt, (b)∫ F dy <strong>and</strong> (c) ∫ F · dr.L L L11.4 By making an appropriate choice <strong>for</strong> the functions P (x, y)<strong>and</strong>Q(x, y) that appearin Green’s theorem in a plane, show that the integral of x − y over the upper halfof the unit circle centred on the origin has the value − 2 . Show the same result3by direct integration in Cartesian coordinates.11.5 Determine the point of intersection P , in the first quadrant, of the two ellipsesx 2a + y22 b =1 <strong>and</strong> x 22 b + y22 a =1. 2Taking b


LINE, SURFACE AND VOLUME INTEGRALSin section 11.3, evaluate the two remaining line integrals <strong>and</strong> hence find the totalarea common to the two ellipses.11.6 By using parameterisations of the <strong>for</strong>m x = a cos n θ <strong>and</strong> y = a sin n θ <strong>for</strong> suitablevalues of n, find the area bounded by the curvesx 2/5 + y 2/5 = a 2/5 <strong>and</strong> x 2/3 + y 2/3 = a 2/3 .11.7 Evaluate the line integral∮[I = y(4x 2 + y 2 ) dx + x(2x 2 +3y 2 ) dy ]Caround the ellipse x 2 /a 2 + y 2 /b 2 =1.11.8 Criticise the following ‘proof’ that π =0.(a) Apply Green’s theorem in a plane to the functions P (x, y) =tan −1 (y/x) <strong>and</strong>Q(x, y) =tan −1 (x/y), taking the region R to be the unit circle centred on theorigin.(b) The RHS of the equality so produced is∫ ∫y − xdx dy,R x 2 + y2 which, either from symmetry considerations or by changing to plane polarcoordinates, can be shown to have zero value.(c) In the LHS of the equality, set x =cosθ <strong>and</strong> y =sinθ, yielding P (θ) =θ<strong>and</strong> Q(θ) =π/2 − θ. The line integral becomes∫ 2π0[( π)]2 − θ cos θ − θ sin θ dθ,which has the value 2π.(d) Thus 2π = 0 <strong>and</strong> the stated result follows.11.9 A single-turn coil C of arbitrary shape is placed in a magnetic field B <strong>and</strong> carriesa current I. Show that the couple acting upon the coil can be written as∫∫M = I (B · r) dr − I B(r · dr).CFor a planar rectangular coil of sides 2a <strong>and</strong> 2b placed with its plane vertical<strong>and</strong> at an angle φ to a uni<strong>for</strong>m horizontal field B, show that M is, as expected,4abBI cos φ k.11.10 Find the vector area S of the part of the curved surface of the hyperboloid ofrevolutionx 2a − y2 + z 2=12 b 2that lies in the region z ≥ 0<strong>and</strong>a ≤ x ≤ λa.11.11 An axially symmetric solid body with its axis AB vertical is immersed in anincompressible fluid of density ρ 0 . Use the following method to show that,whatever the shape of the body, <strong>for</strong> ρ = ρ(z) in cylindrical polars the Archimedeanupthrust is, as expected, ρ 0 gV, whereV is the volume of the body.Express the vertical component of the resultant <strong>for</strong>ce on the body, − ∫ pdS,where p is the pressure, in terms of an integral; note that p = −ρ 0 gz <strong>and</strong> that <strong>for</strong>an annular surface element of width dl, n · n z dl = −dρ. Integrate by parts <strong>and</strong>use the fact that ρ(z A )=ρ(z B )=0.410C


11.10 EXERCISES11.12 Show that the expression below is equal to the solid angle subtended by arectangular aperture, of sides 2a <strong>and</strong> 2b, at a point on the normal through itscentre, <strong>and</strong> at a distance c from the aperture:∫ bacΩ=4dy.0 (y 2 + c 2 )(y 2 + c 2 + a 2 )1/2By setting y =(a 2 + c 2 ) 1/2 tan φ, change this integral into the <strong>for</strong>m∫ φ14ac cos φ0 c 2 + a 2 sin 2 φ dφ,where tan φ 1 = b/(a 2 + c 2 ) 1/2 , <strong>and</strong> hence show that[]Ω=4tan −1 ab.c(a 2 + b 2 + c 2 ) 1/211.13 A vector field a is given by −zxr −3 i−zyr −3 j+(x 2 +y 2 )r −3 k,wherer 2 = x 2 +y 2 +z 2 .Establish that the field is conservative (a) by showing that ∇×a = 0, <strong>and</strong>(b)byconstructing its potential function φ.11.14 A vector field a is given by (z 2 +2xy) i +(x 2 +2yz) j +(y 2 +2zx) k. Show thata is conservative <strong>and</strong> that the line integral ∫ a · dr along any line joining (1, 1, 1)<strong>and</strong> (1, 2, 2) has the value 11.11.15 A <strong>for</strong>ce F(r) acts on a particle at r. In which of the following cases can F berepresented in terms of a potential? Where it can, find the potential.[] )2(x − y)(a) F = F 0 i − j − r exp(− r2;a 2a 2(b) F = F []0zk + (x2 + y 2 − a 2 )r expaa 2(c) F = F 0[k +]a(r × k).r 2(− r2a 2 );11.16 One of Maxwell’s electromagnetic equations states that all magnetic fields Bare solenoidal (i.e. ∇ · B = 0). Determine whether each of the following vectorscould represent a real magnetic field; where it could, try to find a suitable vectorpotential A, i.e. such that B = ∇×A. (Hint: seek a vector potential that is parallelto ∇×B.):B 0 b(a)r [(x − y)z i +(x − y)z j 3 +(x2 − y 2 ) k] in Cartesians with r 2 = x 2 +y 2 +z 2 ;B 0 b 3(b) [cos θ cos φ êr 3r − sin θ cos φ ê θ +sin2θ sin φ ê φ ] in spherical polars;[](c) B 0 b 2 zρ1(b 2 + z 2 ) 2 êρ +b 2 + z 2 êz in cylindrical polars.11.17 The vector field f has components yi−xj+k <strong>and</strong> γ is a curve given parametricallybyr =(a − c + c cos θ)i +(b + c sin θ)j + c 2 θk, 0 ≤ θ ≤ 2π.Describe the shape of the path γ <strong>and</strong> show that the line integral ∫ f · dr vanishes.γDoes this result imply that f is a conservative field?11.18 A vector field a = f(r)r is spherically symmetric <strong>and</strong> everywhere directed awayfrom the origin. Show that a is irrotational, but that it is also solenoidal only iff(r) isofthe<strong>for</strong>mAr −3 .411


LINE, SURFACE AND VOLUME INTEGRALS11.19 Evaluate the surface integral ∫ r·dS, wherer is the position vector, over that partof the surface z = a 2 − x 2 − y 2 <strong>for</strong> which z ≥ 0, by each of the following methods.(a) Parameterise the surface as x = a sin θ cos φ, y = a sin θ sin φ, z = a 2 cos 2 θ,<strong>and</strong> show thatr · dS = a 4 (2 sin 3 θ cos θ +cos 3 θ sin θ) dθ dφ.(b) Apply the divergence theorem to the volume bounded by the surface <strong>and</strong>the plane z =0.11.20 Obtain an expression <strong>for</strong> the value φ P at a point P of a scalar function φ thatsatisfies ∇ 2 φ = 0, in terms of its value <strong>and</strong> normal derivative on a surface S thatencloses it, by proceeding as follows.(a) In Green’s second theorem, take ψ at any particular point Q as 1/r, whereris the distance of Q from P . Show that ∇ 2 ψ = 0, except at r =0.(b) Apply the result to the doubly connected region bounded by S <strong>and</strong> a smallsphere Σ of radius δ centred on P.(c) Apply the divergence theorem to show that the surface integral over Σinvolving 1/δ vanishes, <strong>and</strong> prove that the term involving 1/δ 2 has the value4πφ P .(d) Conclude thatφ P = − 1 ∫φ ∂ ( ) 1dS + 1 ∫4π S ∂n r 4π S1 ∂φr ∂n dS.This important result shows that the value at a point P of a function φthat satisfies ∇ 2 φ = 0 everywhere within a closed surface S that encloses Pmay be expressed entirely in terms of its value <strong>and</strong> normal derivative on S.This matter is taken up more generally in connection with Green’s functionsin chapter 21 <strong>and</strong> in connection with functions of a complex variable insection 24.10.11.21 Use result (11.21), together with an appropriately chosen scalar function φ, toprove that the position vector ¯r of the centre of mass of an arbitrarily shapedbody of volume V <strong>and</strong> uni<strong>for</strong>m density can be written¯r = 1 ∮12Vr2 dS.S11.22 A rigid body of volume V <strong>and</strong> surface S rotates with angular velocity ω. Show thatω = − 1 ∮u × dS,2V Swhere u(x) is the velocity of the point x on the surface S.11.23 Demonstrate the validity of the divergence theorem:(a) by calculating the flux of the vectorαrF =(r 2 + a 2 ) 3/2through the spherical surface |r| = √ 3a;(b) by showing that3αa 2∇ · F =(r 2 + a 2 ) 5/2<strong>and</strong> evaluating the volume integral of ∇ · F over the interior of the sphere|r| = √ 3a. The substitution r = a tan θ will prove useful in carrying out theintegration.412


11.10 EXERCISES11.24 Prove equation (11.22) <strong>and</strong>, by taking b = zx 2 i + zy 2 j +(x 2 − y 2 )k, show that thetwo integrals∫∫I = x 2 dV <strong>and</strong> J = cos 2 θ sin 3 θ cos 2 φdθdφ,both taken over the unit sphere, must have the same value. Evaluate both directlyto show that the common value is 4π/15.11.25 In a uni<strong>for</strong>m conducting medium with unit relative permittivity, charge density ρ,current density J, electric field E <strong>and</strong> magnetic field B, Maxwell’s electromagneticequations take the <strong>for</strong>m (with µ 0 ɛ 0 = c −2 )(i) ∇ · B = 0, (ii) ∇ · E = ρ/ɛ 0 ,(iii) ∇×E + Ḃ = 0, (iv) ∇×B − (Ė/c2 )=µ 0 J.The density of stored energy in the medium is given by 1 (ɛ 2 0E 2 + µ −10 B2 ). Showthat the rate of change of the total stored energy in a volume V is equal to∫− J · E dV − 1 ∮(E × B) · dS,µ 0Vwhere S is the surface bounding V .[ The first integral gives the ohmic heating loss, whilst the second gives theelectromagnetic energy flux out of the bounding surface. The vector µ −10 (E × B)is known as the Poynting vector. ]11.26 A vector field F is defined in cylindrical polar coordinates ρ, θ, z by( )x cos λz y cos λzF = F 0 i + j +(sinλz)k ≡ F 0ρa aa (cos λz)e ρ + F 0 (sin λz)k,where i, j <strong>and</strong> k are the unit vectors along the Cartesian axes <strong>and</strong> e ρ is the unitvector (x/ρ)i +(y/ρ)j.(a) Calculate, as a surface integral, the flux of F through the closed surfacebounded by the cylinders ρ = a <strong>and</strong> ρ =2a <strong>and</strong> the planes z = ±aπ/2.(b) Evaluate the same integral using the divergence theorem.11.27 The vector field F is given byF =(3x 2 yz + y 3 z + xe −x )i +(3xy 2 z + x 3 z + ye x )j +(x 3 y + y 3 x + xy 2 z 2 )k.Calculate (a) directly, <strong>and</strong> (b) by using Stokes’ theorem the value of the lineintegral ∫ F · dr, whereL is the (three-dimensional) closed contour OABCDEOLdefined by the successive vertices (0, 0, 0), (1, 0, 0), (1, 0, 1), (1, 1, 1), (1, 1, 0), (0, 1, 0),(0, 0, 0).11.28 A vector <strong>for</strong>ce field F is defined in Cartesian coordinates by[( y3F = F 03a + y ) ( xy23 a exy/a2 +1 i + + x + y )e xy/a2 j + z ]a 3 aa exy/a2 k .Use Stokes’ theorem to calculate∮F · dr,Lwhere L is the perimeter of the rectangle ABCD given by A =(0, 1, 0), B =(1, 1, 0),C =(1, 3, 0) <strong>and</strong> D =(0, 3, 0).S413


LINE, SURFACE AND VOLUME INTEGRALS11.11 Hints <strong>and</strong> answers11.1 Show that ∇×F = 0. The potential φ F (r) =x 2 z + y 2 z 2 − z.11.3 (a) c 3 ln 2 i +2j +(3c/2)k; (b)(−3c 4 /8)i − c j − (c 2 ln 2)k; (c)c 4 ln 2 − c.11.5 For P , x = y = ab/(a 2 + b 2 ) 1/2 . The relevant limits are 0 ≤ θ 1 ≤ tan −1 (b/a) <strong>and</strong>tan −1 (a/b) ≤ θ 2 ≤ π/2. The total common area is 4ab tan −1 (b/a).11.7 Show that, in the notation of section 11.3, ∂Q/∂x − ∂P/∂y =2x 2 ; I = πa 3 b/2.11.9 M = I ∫ r × (dr × B). Show that the horizontal sides in the first term <strong>and</strong> theCwhole of the second term contribute nothing to the couple.11.11 Note that, if ˆn is the outward normal to the surface, ˆn z · ˆn dl is equal to −dρ.11.13 (b) φ = c + z/r.11.15 (a) Yes, F 0 (x − y)exp(−r 2 /a 2 ); (b) yes, −F 0 [(x 2 + y 2 )/(2a)] exp(−r 2 /a 2 );(c) no, ∇×F ≠ 0.11.17 A spiral of radius c with its axis parallel to the z-direction <strong>and</strong> passing through(a, b). The pitch of the spiral is 2πc 2 . No, because (i) γ is not a closed loop <strong>and</strong>(ii) the line integral must be zero <strong>for</strong> every closed loop, not just <strong>for</strong> a particularone. In fact ∇×f = −2k ≠ 0 shows that f is not conservative.11.19 (a) dS =(2a 3 cos θ sin 2 θ cos φ i +2a 3 cos θ sin 2 θ sin φ j + a 2 cos θ sin θ k) dθ dφ.(b) ∇ · r = 3; over the plane z =0,r · dS =0.The necessarily common value is 3πa 4 /2.11.21 Write r as ∇( 1 2 r2 ).11.23 The answer is 3 √ 3πα/2 in each case.11.25 Identify the expression <strong>for</strong> ∇ · (E × B) <strong>and</strong> use the divergence theorem.11.27 (a) The successive contributions to the integral are:1 − 2e −1 , 0, 2+ 1 2 e, − 7 3 , −1+2e−1 , − 1 2 .(b) ∇×F =2xyz 2 i − y 2 z 2 j + ye x k. Show that the contour is equivalent to thesum of two plane square contours in the planes z =0<strong>and</strong>x = 1, the latter beingtraversed in the negative sense. Integral = 1 (3e − 5).6414


12Fourier seriesWe have already discussed, in chapter 4, how complicated functions may beexpressed as power series. However, this is not the only way in which a functionmay be represented as a series, <strong>and</strong> the subject of this chapter is the expressionof functions as a sum of sine <strong>and</strong> cosine terms. Such a representation is called aFourier series. Unlike Taylor series, a Fourier series can describe functions that arenot everywhere continuous <strong>and</strong>/or differentiable. There are also other advantagesin using trigonometric terms. They are easy to differentiate <strong>and</strong> integrate, theirmoduli are easily taken <strong>and</strong> each term contains only one characteristic frequency.This last point is important because, as we shall see later, Fourier series are oftenused to represent the response of a system to a periodic input, <strong>and</strong> this responseoften depends directly on the frequency content of the input. Fourier series areused in a wide variety of such physical situations, including the vibrations of afinite string, the scattering of light by a diffraction grating <strong>and</strong> the transmissionof an input signal by an electronic circuit.12.1 The Dirichlet conditionsWe have already mentioned that Fourier series may be used to represent somefunctions <strong>for</strong> which a Taylor series expansion is not possible. The particularconditions that a function f(x) must fulfil in order that it may be exp<strong>and</strong>ed as aFourier series are known as the Dirichlet conditions, <strong>and</strong> may be summarised bythe following four points:(i) the function must be periodic;(ii) it must be single-valued <strong>and</strong> continuous, except possibly at a finite numberof finite discontinuities;(iii) it must have only a finite number of maxima <strong>and</strong> minima within oneperiod;(iv) the integral over one period of |f(x)| must converge.415


FOURIER SERIESf(x)xLLFigure 12.1 An example of a function that may be represented as a Fourierseries without modification.If the above conditions are satisfied then the Fourier series converges to f(x)at all points where f(x) is continuous. The convergence of the Fourier seriesat points of discontinuity is discussed in section 12.4. The last three Dirichletconditions are almost always met in real applications, but not all functions areperiodic <strong>and</strong> hence do not fulfil the first condition. It may be possible, however,to represent a non-periodic function as a Fourier series by manipulation of thefunction into a periodic <strong>for</strong>m. This is discussed in section 12.5. An example ofa function that may, without modification, be represented as a Fourier series isshown in figure 12.1.We have stated without proof that any function that satisfies the Dirichletconditions may be represented as a Fourier series. Let us now show why this isa plausible statement. We require that any reasonable function (one that satisfiesthe Dirichlet conditions) can be expressed as a linear sum of sine <strong>and</strong> cosineterms. We first note that we cannot use just a sum of sine terms since sine, beingan odd function (i.e. a function <strong>for</strong> which f(−x) =−f(x)), cannot represent evenfunctions (i.e. functions <strong>for</strong> which f(−x) =f(x)). This is obvious when we tryto express a function f(x) that takes a non-zero value at x = 0. Clearly, sincesin nx = 0 <strong>for</strong> all values of n, we cannot represent f(x) atx = 0 by a sine series.Similarly odd functions cannot be represented by a cosine series since cosine isan even function. Nevertheless, it is possible to represent all odd functions by asine series <strong>and</strong> all even functions by a cosine series. Now, since all functions maybewrittenasthesumofanodd<strong>and</strong>anevenpart,f(x) = 1 2 [f(x)+f(−x)] + 1 2[f(x) − f(−x)]= f even (x)+f odd (x),416


12.2 THE FOURIER COEFFICIENTSwe can write any function as the sum of a sine series <strong>and</strong> a cosine series.All the terms of a Fourier series are mutually orthogonal, i.e. the integrals, overone period, of the product of any two terms have the following properties:∫ x0+L( ) ( )2πrx 2πpxsin cos dx = 0 <strong>for</strong> all r <strong>and</strong> p, (12.1)L Lx 0∫ x0+Lx 0∫ x0+Lx 0( 2πrxcosL( 2πrxsinL)cos)sin⎧( ) 2πpx⎨L <strong>for</strong> r = p =0,1dx =L ⎩2L <strong>for</strong> r = p>0,0 <strong>for</strong> r ≠ p,⎧( ) 2πpx⎨0 <strong>for</strong> r = p =0,1dx =L ⎩2L <strong>for</strong> r = p>0,0 <strong>for</strong> r ≠ p,(12.2)(12.3)where r <strong>and</strong> p are integers greater than or equal to zero; these <strong>for</strong>mulae are easilyderived. A full discussion of why it is possible to exp<strong>and</strong> a function as a sum ofmutually orthogonal functions is given in chapter 17.The Fourier series expansion of the function f(x) is conventionally writtenf(x) = a ∞02 + ∑[a r cosr=1( 2πrxL)+ b r sin( 2πrxL)], (12.4)where a 0 ,a r ,b r are constants called the Fourier coefficients. These coefficients areanalogous to those in a power series expansion <strong>and</strong> the determination of theirnumerical values is the essential step in writing a function as a Fourier series.This chapter continues with a discussion of how to find the Fourier coefficients<strong>for</strong> particular functions. We then discuss simplifications to the general Fourierseries that may save considerable ef<strong>for</strong>t in calculations. This is followed by thealternative representation of a function as a complex Fourier series, <strong>and</strong> weconclude with a discussion of Parseval’s theorem.12.2 The Fourier coefficientsWe have indicated that a series that satisfies the Dirichlet conditions may bewritten in the <strong>for</strong>m (12.4). We now consider how to find the Fourier coefficients<strong>for</strong> any particular function. For a periodic function f(x) ofperiodL we will findthat the Fourier coefficients are given bya r = 2 ∫ x0+L( ) 2πrxf(x)cos dx, (12.5)L x 0Lb r = 2 ∫ x0+L( ) 2πrxf(x)sin dx, (12.6)L x 0Lwhere x 0 is arbitrary but is often taken as 0 or −L/2. The apparently arbitraryfactor 1 2which appears in the a 0 term in (12.4) is included so that (12.5) may417


FOURIER SERIESapply <strong>for</strong> r = 0 as well as r>0. The relations (12.5) <strong>and</strong> (12.6) may be derivedas follows.Suppose the Fourier series expansion of f(x) can be written as in (12.4),f(x) = a ∞ 02 + ∑[a r cosr=1( 2πrxL)+ b r sin( 2πrxL)].Then, multiplying by cos(2πpx/L), integrating over one full period in x <strong>and</strong>changing the order of the summation <strong>and</strong> integration, we get∫ x0+Lx 0( ) 2πpxf(x)cosLdx = a 02∫ x0+L( ) 2πpxcos dxx 0L∞∑∫ x0+L( 2πrx+ a r cosr=1x 0L∞∑∫ x0+L( 2πrx+ b r sinLr=1x 0)cos)cos( ) 2πpxdxL( ) 2πpxdx.L(12.7)We can now find the Fourier coefficients by considering (12.7) as p takes differentvalues. Using the orthogonality conditions (12.1)–(12.3) of the previous section,we find that when p = 0 (12.7) becomes∫ x0+Lf(x)dx = a 0x 02 L.When p ≠ 0 the only non-vanishing term on the RHS of (12.7) occurs whenr = p, <strong>and</strong>so∫ x0+L( ) 2πrxf(x)cos dx = a rL 2 L.x 0The other Fourier coefficients b r may be found by repeating the above processbut multiplying by sin(2πpx/L) instead of cos(2πpx/L) (see exercise 12.2).◮Express the square-wave function illustrated in figure 12.2 as a Fourier series.Physically this might represent the input to an electrical circuit that switches between ahigh <strong>and</strong> a low state with time period T . The square wave may be represented by{−1 <strong>for</strong> − 1f(t) =T ≤ t


12.3 SYMMETRY CONSIDERATIONSf(t)1− T 20 T2t−1Figure 12.2A square-wave function.following section). To evaluate the coefficients in the sine series we use (12.6). Henceb r = 2 ∫ T/2( ) 2πrtf(t)sin dtT −T/2 T= 4 ∫ T/2( ) 2πrtsin dtT 0 T= 2 πr [1 − (−1)r ] .Thus the sine coefficients are zero if r is even <strong>and</strong> equal to 4/(πr) ifr is odd. Hence theFourier series <strong>for</strong> the square-wave function may be written as()f(t) = 4 sin 3ωt sin 5ωtsin ωt + + + ···π3 5, (12.8)where ω =2π/T is called the angular frequency. ◭12.3 Symmetry considerationsThe example in the previous section employed the useful property that since thefunction to be represented was odd, all the cosine terms of the Fourier series wereabsent. It is often the case that the function we wish to express as a Fourier serieshas a particular symmetry, which we can exploit to reduce the calculational labourof evaluating Fourier coefficients. Functions that are symmetric or antisymmetricabout the origin (i.e. even <strong>and</strong> odd functions respectively) admit particularlyuseful simplifications. Functions that are odd in x have no cosine terms (seesection 12.1) <strong>and</strong> all the a-coefficients are equal to zero. Similarly, functions thatare even in x have no sine terms <strong>and</strong> all the b-coefficients are zero. Since theFourier series of odd or even functions contain only half the coefficients required<strong>for</strong> a general periodic function, there is a considerable reduction in the algebraneeded to find a Fourier series.The consequences of symmetry or antisymmetry of the function about thequarter period (i.e. about L/4) are a little less obvious. Furthermore, the results419


FOURIER SERIESare not used as often as those above <strong>and</strong> the remainder of this section can beomitted on a first reading without loss of continuity. The following argumentgives the required results.Suppose that f(x) has even or odd symmetry about L/4, i.e. f(L/4 − x) =±f(x − L/4). For convenience, we make the substitution s = x − L/4 <strong>and</strong> hencef(−s) =±f(s). We can now see thatb r = 2 ∫ x0+L( 2πrsf(s)sinL x 0L+ πr )ds,2where the limits of integration have been left unaltered since f is, of course,periodic in s as well as in x. If we use the expansion( 2πrssinL+ πr ) ( ) 2πrs( πr) ( 2πrs=sin cos +cos2L 2L)sin( πr),2we can immediately see that the trigonometric part of the integr<strong>and</strong> is an oddfunction of s if r is even <strong>and</strong> an even function of s if r is odd. Hence if f(s) iseven <strong>and</strong> r is even then the integral is zero, <strong>and</strong> if f(s) is odd <strong>and</strong> r is odd thenthe integral is zero. Similar results can be derived <strong>for</strong> the Fourier a-coefficients<strong>and</strong> we conclude that(i) if f(x) is even about L/4 thena 2r+1 = 0 <strong>and</strong> b 2r =0,(ii) if f(x) is odd about L/4 thena 2r = 0 <strong>and</strong> b 2r+1 =0.All the above results follow automatically when the Fourier coefficients areevaluated in any particular case, but prior knowledge of them will often enablesome coefficients to be set equal to zero on inspection <strong>and</strong> so substantially reducethe computational labour. As an example, the square-wave function shown infigure 12.2 is (i) an odd function of t, sothatalla r = 0, <strong>and</strong> (ii) even about thepoint t = T/4, so that b 2r = 0. Thus we can say immediately that only sine termsof odd harmonics will be present <strong>and</strong> there<strong>for</strong>e will need to be calculated; this isconfirmed in the expansion (12.8).12.4 Discontinuous functionsThe Fourier series expansion usually works well <strong>for</strong> functions that are discontinuousin the required range. However, the series itself does not produce adiscontinuous function <strong>and</strong> we state without proof that the value of the exp<strong>and</strong>edf(x) at a discontinuity will be half-way between the upper <strong>and</strong> lowervalues. Expressing this more mathematically, at a point of finite discontinuity, x d ,the Fourier series converges to12 lim[f(xd + ɛ)+f(x d − ɛ)].ɛ→0At a discontinuity, the Fourier series representation of the function will overshootits value. Although as more terms are included the overshoot moves in position420


12.4 DISCONTINUOUS FUNCTIONS(a)1(b)1− T 2T2− T 2T2−1−1(c)1(d)1δ− T 2T2− T 2T2−1−1Figure 12.3 The convergence of a Fourier series expansion of a square-wavefunction, including (a) one term, (b) two terms, (c) three terms <strong>and</strong> (d) 20terms. The overshoot δ is shown in (d).arbitrarily close to the discontinuity, it never disappears even in the limit of aninfinite number of terms. This behaviour is known as Gibbs’ phenomenon. Afulldiscussion is not pursued here but suffice it to say that the size of the overshootis proportional to the magnitude of the discontinuity.◮Find the value to which the Fourier series of the square-wave function discussed in section12.2 converges at t =0.It can be seen that the function is discontinuous at t = 0 <strong>and</strong>, by the above rule, we expectthe series to converge to a value half-way between the upper <strong>and</strong> lower values, in otherwords to converge to zero in this case. Considering the Fourier series of this function,(12.8), we see that all the terms are zero <strong>and</strong> hence the Fourier series converges to zero asexpected. The Gibbs phenomenon <strong>for</strong> the square-wave function is shown in figure 12.3. ◭421


FOURIER SERIES(a)0L(b)0L2L(c)0L2L(d)0L2LFigure 12.4Possible periodic extensions of a function.12.5 Non-periodic functionsWe have already mentioned that a Fourier representation may sometimes be used<strong>for</strong> non-periodic functions. If we wish to find the Fourier series of a non-periodicfunction only within a fixed range then we may continue the function outside therange so as to make it periodic. The Fourier series of this periodic function wouldthen correctly represent the non-periodic function in the desired range. Since weare often at liberty to extend the function in a number of ways, we can sometimesmake it odd or even <strong>and</strong> so reduce the calculation required. Figure 12.4(b) showsthe simplest extension to the function shown in figure 12.4(a). However, thisextension has no particular symmetry. Figures 12.4(c), (d) show extensions as odd<strong>and</strong> even functions respectively with the benefit that only sine or cosine termsappear in the resulting Fourier series. We note that these last two extensions givea function of period 2L.In view of the result of section 12.4, it must be added that the continuationmust not be discontinuous at the end-points of the interval of interest; if it isthe series will not converge to the required value there. This requirement thatthe series converges appropriately may reduce the choice of continuations. Thisis discussed further at the end of the following example.◮Find the Fourier series of f(x) =x 2 <strong>for</strong> 0


12.5 NON-PERIODIC FUNCTIONSf(x) =x 2−2 0 2xLFigure 12.5f(x) =x 2 ,0


FOURIER SERIESconverge to the correct values of f(x) =±4 atx = ±2; it converges, instead, tozero, the average of the values at the two ends of the range.12.6 Integration <strong>and</strong> differentiationIt is sometimes possible to find the Fourier series of a function by integration ordifferentiation of another Fourier series. If the Fourier series of f(x) isintegratedterm by term then the resulting Fourier series converges to the integral of f(x).Clearly, when integrating in such a way there is a constant of integration that mustbe found. If f(x) is a continuous function of x <strong>for</strong> all x <strong>and</strong> f(x) is also periodicthen the Fourier series that results from differentiating term by term converges tof ′ (x), provided that f ′ (x) itself satisfies the Dirichlet conditions. These propertiesof Fourier series may be useful in calculating complicated Fourier series, sincesimple Fourier series may easily be evaluated (or found from st<strong>and</strong>ard tables)<strong>and</strong> often the more complicated series can then be built up by integration <strong>and</strong>/ordifferentiation.◮Find the Fourier series of f(x) =x 3 <strong>for</strong> 0


12.7 COMPLEX FOURIER SERIESwhere the Fourier coefficients are given byc r = 1 ∫ x0+L(f(x)exp − 2πirx )dx. (12.10)LLx 0This relation can be derived, in a similar manner to that of section 12.2, by multiplying(12.9) by exp(−2πipx/L) be<strong>for</strong>e integrating <strong>and</strong> using the orthogonalityrelation∫ x0+L(exp − 2πipx ) ( ) {2πirx L <strong>for</strong> r = p,exp dx =LL0 <strong>for</strong> r ≠ p.x 0The complex Fourier coefficients in (12.9) have the following relations to the realFourier coefficients:c r = 1 2 (a r − ib r ),c −r = 1 2 (a r + ib r ).(12.11)Note that if f(x) isrealthenc −r = c ∗ r , where the asterisk represents complexconjugation.◮Find a complex Fourier series <strong>for</strong> f(x) =x in the range −2


FOURIER SERIES12.8 Parseval’s theoremParseval’s theorem gives a useful way of relating the Fourier coefficients to thefunction that they describe. Essentially a conservation law, it states that∫1 x0+L∞∑|f(x)| 2 dx = |c r | 2L x 0r=−∞= ( 12 a ) 20 +12∞∑(a 2 r + b 2 r ). (12.13)In a more memorable <strong>for</strong>m, this says that the sum of the moduli squared ofthe complex Fourier coefficients is equal to the average value of |f(x)| 2 over oneperiod. Parseval’s theorem can be proved straight<strong>for</strong>wardly by writing f(x) asa Fourier series <strong>and</strong> evaluating the required integral, but the algebra is messy.There<strong>for</strong>e, we shall use an alternative method, <strong>for</strong> which the algebra is simple<strong>and</strong> which in fact leads to a more general <strong>for</strong>m of the theorem.Let us consider two functions f(x) <strong>and</strong> g(x), which are (or can be made)periodic with period L <strong>and</strong> which have Fourier series (expressed in complex<strong>for</strong>m)∞∑( ) 2πirxf(x) = c r exp ,Lr=−∞∞∑( ) 2πirxg(x) = γ r exp ,Lr=−∞where c r <strong>and</strong> γ r are the complex Fourier coefficients of f(x) <strong>and</strong>g(x) respectively.Thus∞∑( ) 2πirxf(x)g ∗ (x) = c r g ∗ (x)exp .Lr=−∞Integrating this equation with respect to x over the interval (x 0 ,x 0 + L) <strong>and</strong>dividing by L, we find∫1 x0+L∞∑∫f(x)g ∗ 1 x0+L( ) 2πirx(x) dx = c r g ∗ (x)exp dxL x 0Lr=−∞ x 0L∞∑[ ∫ 1 x0+L( ) ] ∗−2πirx= c r g(x)expdxLr=−∞ x 0L∞∑= c r γr ∗ ,r=−∞where the last equality uses (12.10). Finally, if we let g(x) =f(x) then we obtainParseval’s theorem (12.13). This result can be proved in a similar manner using426r=1


12.9 EXERCISESthe sine <strong>and</strong> cosine <strong>for</strong>m of the Fourier series, but the algebra is slightly morecomplicated.Parseval’s theorem is sometimes used to sum series. However, if one is presentedwith a series to sum, it is not usually possible to decide which Fourier seriesshould be used to evaluate it. Rather, useful summations are nearly always foundserendipitously. The following example shows the evaluation of a sum by aFourier series method.◮Using Parseval’s theorem <strong>and</strong> the Fourier series <strong>for</strong> f(x) =x 2 found in section 12.5,calculate the sum ∑ ∞r=1 r−4 .Firstly we find the average value of [f(x)] 2 over the interval −2


FOURIER SERIESbe better <strong>for</strong> numerical evaluation? Relate your answer to the relevant periodiccontinuations.12.7 For the continued functions used in exercise 12.6 <strong>and</strong> the derived correspondingseries, consider (i) their derivatives <strong>and</strong> (ii) their integrals. Do they give meaningfulequations? You will probably find it helpful to sketch all the functions involved.12.8 The function y(x) =x sin x <strong>for</strong> 0 ≤ x ≤ π is to be represented by a Fourier seriesof period 2π that is either even or odd. By sketching the function <strong>and</strong> consideringits derivative, determine which series will have the more rapid convergence. Findthe full expression <strong>for</strong> the better of these two series, showing that the convergence∼ n −3 <strong>and</strong> that alternate terms are missing.12.9 Find the Fourier coefficients in the expansion of f(x) =expx over the range−1


12.9 EXERCISESDeduce the value of the sum S of the series1 − 1 3 3 + 1 5 3 − 1 7 3 + ··· .12.15 Using the result of exercise 12.14, determine, as far as possible by inspection, the<strong>for</strong>ms of the functions of which the following are the Fourier series:(a)(b)cos θ + 1 9 cos 3θ + 1 cos 5θ + ··· ;25sin θ + 1 1sin 3θ + sin 5θ + ··· ;27 125(c)[L 23 − 4L2 cos πxπ 2 L − 1 2πxcos4 L+ 1 3πx···]cos9 L − .(You may find it helpful to first set x = 0 in the quoted result <strong>and</strong> so obtainvalues <strong>for</strong> S o = ∑ (2m +1) −2 <strong>and</strong> other sums derivable from it.)12.16 By finding a cosine Fourier series of period 2 <strong>for</strong> the function f(t) thattakesthe<strong>for</strong>m f(t) =cosh(t − 1) in the range 0 ≤ t ≤ 1, prove that∞∑n=11n 2 π 2 +1 = 1e 2 − 1 .Deduce values <strong>for</strong> the sums ∑ (n 2 π 2 +1) −1 over odd n <strong>and</strong> even n separately.12.17 Find the (real) Fourier series of period 2 <strong>for</strong> f(x) =coshx <strong>and</strong> g(x) =x 2 in therange −1 ≤ x ≤ 1. By integrating the series <strong>for</strong> f(x) twice, prove that∞∑n=1(−1) n+1n 2 π 2 (n 2 π 2 +1) = 1 2( 1sinh 1 − 5 612.18 Express the function f(x) =x 2 as a Fourier sine series in the range 0


FOURIER SERIES12.21 Find the complex Fourier series <strong>for</strong> the periodic function of period 2π defined inthe range −π ≤ x ≤ π by y(x) =coshx. By setting x = 0 prove that∞∑ (−1) nn 2 +1 = 1 ( π)2 sinh π − 1 .n=112.22 The repeating output from an electronic oscillator takes the <strong>for</strong>m of a sine wavef(t) =sint <strong>for</strong> 0 ≤ t ≤ π/2; it then drops instantaneously to zero <strong>and</strong> startsagain. The output is to be represented by a complex Fourier series of the <strong>for</strong>m∞∑c n e 4nti .n=−∞Sketch the function <strong>and</strong> find an expression <strong>for</strong> c n .Verifythatc −n = c ∗ n.Demonstratethat setting t =0<strong>and</strong>t = π/2 produces differing values <strong>for</strong> the sum∞∑ 116n 2 − 1 .n=1Determine the correct value <strong>and</strong> check it using the result of exercise 12.20.12.23 Apply Parseval’s theorem to the series found in the previous exercise <strong>and</strong> soderive a value <strong>for</strong> the sum of the series17(15) + 652 (63) + 1452 (143) + ···+ 16n2 +12 (16n 2 − 1) + ··· .212.24 A string, anchored at x = ±L/2, has a fundamental vibration frequency of 2L/c,where c is the speed of transverse waves on the string. It is pulled aside at itscentre point by a distance y 0 <strong>and</strong> released at time t = 0. Its subsequent motioncan be described by the series∞∑y(x, t) = a n cos nπx nπctcosL L .n=1Find a general expression <strong>for</strong> a n <strong>and</strong> show that only the odd harmonics of thefundamental frequency are present in the sound generated by the released string.By applying Parseval’s theorem, find the sum S of the series ∑ ∞0 (2m +1)−4 .12.25 Show that Parseval’s theorem <strong>for</strong> two real functions whose Fourier expansionshave cosine <strong>and</strong> sine coefficients a n , b n <strong>and</strong> α n , β n takes the <strong>for</strong>m∫1 Lf(x)g ∗ (x) dx = 1 L 04 a 0α 0 + 1 ∞∑(a n α n + b n β n ).2n=1(a) Demonstrate that <strong>for</strong> g(x) =sinmx or cos mx this reduces to the definitionof the Fourier coefficients.(b) Explicitly verify the above result <strong>for</strong> the case in which f(x) =x <strong>and</strong> g(x) isthe square-wave function, both in the interval −1 ≤ x ≤ 1.12.26 An odd function f(x) ofperiod2π is to be approximated by a Fourier sine serieshaving only m terms. The error in this approximation is measured by the squaredeviation∫ [2 πm∑E m = f(x) − b n sin nx]dx.−πn=1By differentiating E m with respect to the coefficients b n , find the values of b n thatminimise E m .430


12.10 HINTS AND ANSWERS0 1 0 1 0 1 0 2 4(a) (b) (c) (d)Figure 12.6 Continuations of exp(−x 2 )in0≤ x ≤ 1 to give: (a) cosine termsonly; (b) sine terms only; (c) period 1; (d) period 2.Sketch the graph of the function f(x), where{−x(π + x) <strong>for</strong> −π ≤ x


FOURIER SERIES12.21 c n =[(−1) n sinh π]/[π(1 + n 2 )]. Having set x = 0, separate out the n =0term<strong>and</strong> note that (−1) n =(−1) −n .12.23 (π 2 − 8)/16.12.25 (b) All a n <strong>and</strong> α n are zero; b n =2(−1) n+1 /(nπ) <strong>and</strong>β n =4/(nπ). You will needthe result quoted in exercise 12.19.432


13Integral trans<strong>for</strong>msIn the previous chapter we encountered the Fourier series representation of aperiodic function in a fixed interval as a superposition of sinusoidal functions. It isoften desirable, however, to obtain such a representation even <strong>for</strong> functions definedover an infinite interval <strong>and</strong> with no particular periodicity. Such a representationis called a Fourier trans<strong>for</strong>m <strong>and</strong> is one of a class of representations called integraltrans<strong>for</strong>ms.We begin by considering Fourier trans<strong>for</strong>ms as a generalisation of Fourierseries. We then go on to discuss the properties of the Fourier trans<strong>for</strong>m <strong>and</strong> itsapplications. In the second part of the chapter we present an analogous discussionof the closely related Laplace trans<strong>for</strong>m.13.1 Fourier trans<strong>for</strong>msThe Fourier trans<strong>for</strong>m provides a representation of functions defined over aninfinite interval <strong>and</strong> having no particular periodicity, in terms of a superpositionof sinusoidal functions. It may thus be considered as a generalisation of theFourier series representation of periodic functions. Since Fourier trans<strong>for</strong>ms areoften used to represent time-varying functions, we shall present much of ourdiscussion in terms of f(t), rather than f(x), although in some spatial examplesf(x) will be the more natural notation <strong>and</strong> we shall use it as appropriate. Our|f(t)| dt is finite.In order to develop the transition from Fourier series to Fourier trans<strong>for</strong>ms, wefirst recall that a function of period T may be represented as a complex Fourierseries, cf. (12.9),only requirement on f(t) will be that ∫ ∞−∞f(t) =∞∑c r e 2πirt/T =r=−∞∞∑c r e iωrt , (13.1)r=−∞where ω r =2πr/T. As the period T tends to infinity, the ‘frequency quantum’433


INTEGRAL TRANSFORMSc(ω)expiωt− 2πT0−1 02πT4πTω r1 2 rFigure 13.1 The relationship between the Fourier terms <strong>for</strong> a function ofperiod T <strong>and</strong> the Fourier integral (the area below the solid line) of thefunction.∆ω =2π/T becomes vanishingly small <strong>and</strong> the spectrum of allowed frequenciesω r becomes a continuum. Thus, the infinite sum of terms in the Fourier seriesbecomes an integral, <strong>and</strong> the coefficients c r become functions of the continuousvariable ω, as follows.We recall, cf. (12.10), that the coefficients c r in (13.1) are given by∫ T/2∫ T/2c r = 1 f(t) e −2πirt/T dt = ∆ω f(t) e −iωrt dt, (13.2)T −T/22π −T/2where we have written the integral in two alternative <strong>for</strong>ms <strong>and</strong>, <strong>for</strong> convenience,made one period run from −T/2to+T/2 rather than from 0 to T . Substitutingfrom (13.2) into (13.1) gives∞∑∫∆ω T/2f(t) =f(u) e −iωru du e iωrt . (13.3)2πr=−∞ −T/2At this stage ω r is still a discrete function of r equal to 2πr/T.The solid points in figure 13.1 are a plot of (say, the real part of) c r e iωrt asa function of r (or equivalently of ω r ) <strong>and</strong> it is clear that (2π/T)c r e iωrt givesthe area of the rth broken-line rectangle. If T tends to ∞ then ∆ω (= 2π/T)becomes infinitesimal, the width of the rectangles tends to zero <strong>and</strong>, from themathematical definition of an integral,∞∑r=−∞In this particular case∆ω2π g(ω r) e iωrt → 12π∫ ∞−∞∫ T/2g(ω r )= f(u) e −iωru du,−T/2434g(ω) e iωt dω.


13.1 FOURIER TRANSFORMS<strong>and</strong> (13.3) becomesf(t) = 1 ∫ ∞ ∫ ∞dω e iωt du f(u) e −iωu . (13.4)2π −∞−∞This result is known as Fourier’s inversion theorem.From it we may define the Fourier trans<strong>for</strong>m of f(t) by<strong>and</strong> its inverse by˜f(ω) =1√2π∫ ∞−∞f(t) = 1 √2π∫ ∞−∞f(t) e −iωt dt, (13.5)˜f(ω) e iωt dω. (13.6)Including the constant 1/ √ 2π in the definition of ˜f(ω) (whose mathematicalexistence as T →∞is assumed here without proof) is clearly arbitrary, the onlyrequirement being that the product of the constants in (13.5) <strong>and</strong> (13.6) shouldequal 1/(2π). Our definition is chosen to be as symmetric as possible.◮ Find the Fourier trans<strong>for</strong>m of the exponential decay function f(t) =0<strong>for</strong> t0).Using the definition (13.5) <strong>and</strong> separating the integral into two parts,∫1 0˜f(ω) = √ (0) e −iωt dt + A ∫ ∞√ e −λt e −iωt dt2π −∞2π 0=0+ √ A [ ] ∞− e−(λ+iω)t2π λ + iω0A= √ ,2π(λ + iω)which is the required trans<strong>for</strong>m. It is clear that the multiplicative constant A does notaffect the <strong>for</strong>m of the trans<strong>for</strong>m, merely its amplitude. This trans<strong>for</strong>m may be verified byresubstitution of the above result into (13.6) to recover f(t), but evaluation of the integralrequires the use of complex-variable contour integration (chapter 24). ◭13.1.1 The uncertainty principleAn important function that appears in many areas of physical science, eitherprecisely or as an approximation to a physical situation, is the Gaussian ornormal distribution. Its Fourier trans<strong>for</strong>m is of importance both in itself <strong>and</strong> alsobecause, when interpreted statistically, it readily illustrates a <strong>for</strong>m of uncertaintyprinciple.435


INTEGRAL TRANSFORMS◮Find the Fourier trans<strong>for</strong>m of the normalised Gaussian distributionf(t) = 1 ( )τ √ 2π exp − t2, −∞


13.1 FOURIER TRANSFORMSis a wavefunction <strong>and</strong> the distribution of the wave intensity in time is given by|f| 2 (also a Gaussian). Similarly, the intensity distribution in frequency is givenby |˜f| 2 . These two distributions have respective root mean square deviations ofτ/ √ 2<strong>and</strong>1/( √ 2τ), giving, after incorporation of the above relations,∆E ∆t = /2 <strong>and</strong> ∆p ∆x = /2.The factors of 1/2 that appear are specific to the Gaussian <strong>for</strong>m, but anydistribution f(t) produces <strong>for</strong> the product ∆E∆t a quantity λ in which λ isstrictly positive (in fact, the Gaussian value of 1/2 is the minimum possible).13.1.2 Fraunhofer diffractionWe take our final example of the Fourier trans<strong>for</strong>m from the field of optics. Thepattern of transmitted light produced by a partially opaque (or phase-changing)object upon which a coherent beam of radiation falls is called a diffraction pattern<strong>and</strong>, in particular, when the cross-section of the object is small compared withthe distance at which the light is observed the pattern is known as a Fraunhoferdiffraction pattern.We will consider only the case in which the light is monochromatic withwavelength λ. The direction of the incident beam of light can then be describedby the wave vector k; the magnitude of this vector is given by the wave numberk =2π/λ of the light. The essential quantity in a Fraunhofer diffraction patternis the dependence of the observed amplitude (<strong>and</strong> hence intensity) on the angle θbetween the viewing direction k ′ <strong>and</strong> the direction k of the incident beam. Thisis entirely determined by the spatial distribution of the amplitude <strong>and</strong> phase ofthe light at the object, the transmitted intensity in a particular direction k ′ beingdetermined by the corresponding Fourier component of this spatial distribution.As an example, we take as an object a simple two-dimensional screen of width2Y on which light of wave number k is incident normally; see figure 13.2. Wesuppose that at the position (0,y) the amplitude of the transmitted light is f(y)per unit length in the y-direction (f(y) may be complex). The function f(y) iscalled an aperture function. Both the screen <strong>and</strong> beam are assumed infinite in thez-direction.Denoting the unit vectors in the x- <strong>and</strong>y- directions by i <strong>and</strong> j respectively,the total light amplitude at a position r 0 = x 0 i + y 0 j, with x 0 > 0, will be thesuperposition of all the (Huyghens’) wavelets originating from the various partsof the screen. For large r 0 (= |r 0 |), these can be treated as plane waves to give §A(r 0 )=∫ Y−Yf(y) exp[ik ′ · (r 0 − yj)]dy. (13.8)|r 0 − yj|§ This is the approach first used by Fresnel. For simplicity we have omitted from the integral amultiplicative inclination factor that depends on angle θ <strong>and</strong> decreases as θ increases.437


INTEGRAL TRANSFORMSYyk ′k0θx−YFigure 13.2 Diffraction grating of width 2Y with light of wavelength 2π/kbeing diffracted through an angle θ.The factor exp[ik ′ · (r 0 − yj)] represents the phase change undergone by the lightin travelling from the point yj on the screen to the point r 0 , <strong>and</strong> the denominatorrepresents the reduction in amplitude with distance. (Recall that the system isinfinite in the z-direction <strong>and</strong> so the ‘spreading’ is effectively in two dimensionsonly.)If the medium is the same on both sides of the screen then k ′ = k cos θ i+k sin θ j,<strong>and</strong> if r 0 ≫ Y then expression (13.8) can be approximated by∫A(r 0 )= exp(ik′ · r 0 ) ∞f(y) exp(−iky sin θ) dy. (13.9)r 0 −∞We have used that f(y) =0<strong>for</strong>|y| >Y to extend the integral to infinite limits.The intensity in the direction θ is then given byI(θ) =|A| 2 = 2πr2 |˜f(q)| 2 , (13.10)0where q = k sin θ.◮Evaluate I(θ) <strong>for</strong> an aperture consisting of two long slits each of width 2b whose centresare separated by a distance 2a, a>b; the slits are illuminated by light of wavelength λ.The aperture function is plotted in figure 13.3. We first need to find ˜f(q):∫1 −a+b˜f(q) = √ e −iqx dx + 1 ∫ a+b√ e −iqx dx2π 2π−a−b] −a+ba−b] a+b= √ 1 [− e−iqx+ 1 [√ − e−iqx2π iq−a−b 2π iqa−b= −1iq √ [e −iq(−a+b) − e −iq(−a−b) + e −iq(a+b) − e −iq(a−b)] .2π438


13.1 FOURIER TRANSFORMSf(y)1−a−a − b −a + ba − baa + bxFigure 13.3The aperture function f(y) <strong>for</strong> two wide slits.After some manipulation we obtain4cosqa sin qb˜f(q) =q √ .2πNow applying (13.10), <strong>and</strong> remembering that q =(2π sin θ)/λ, we findI(θ) = 16 cos2 qa sin 2 qb,q 2 r2 0where r 0 is the distance from the centre of the aperture. ◭13.1.3 The Dirac δ-functionBe<strong>for</strong>e going on to consider further properties of Fourier trans<strong>for</strong>ms we make adigression to discuss the Dirac δ-function <strong>and</strong> its relation to Fourier trans<strong>for</strong>ms.The δ-function is different from most functions encountered in the physicalsciences but we will see that a rigorous mathematical definition exists; the utilityof the δ-function will be demonstrated throughout the remainder of this chapter.It can be visualised as a very sharp narrow pulse (in space, time, density, etc.)which produces an integrated effect having a definite magnitude. The <strong>for</strong>malproperties of the δ-function may be summarised as follows.The Dirac δ-function has the property thatδ(t) =0 <strong>for</strong>t ≠0, (13.11)but its fundamental defining property is∫f(t)δ(t − a) dt = f(a), (13.12)provided the range of integration includes the point t = a; otherwise the integral439


INTEGRAL TRANSFORMSequals zero. This leads immediately to two further useful results:<strong>and</strong>∫ b−aδ(t) dt = 1 <strong>for</strong> all a, b > 0 (13.13)∫δ(t − a) dt =1, (13.14)provided the range of integration includes t = a.Equation (13.12) can be used to derive further useful properties of the Diracδ-function:δ(t) =δ(−t), (13.15)δ(at) = 1 δ(t), (13.16)|a|tδ(t) =0. (13.17)◮Prove that δ(bt) =δ(t)/|b|.Let us first consider the case where b>0. It follows that∫ ∞∫ ∞( ) t′f(t)δ(bt) dt = f δ(t ′ ) dt′−∞−∞ b b = 1 b f(0) = 1 ∫ ∞f(t)δ(t) dt,b −∞where we have made the substitution t ′ = bt. Butf(t) is arbitrary <strong>and</strong> so we immediatelysee that δ(bt) =δ(t)/b = δ(t)/|b| <strong>for</strong> b>0.Now consider the case where b = −c


13.1 FOURIER TRANSFORMSThe derivative of the delta function, δ ′ (t), is defined by∫ ∞−∞[ ] ∞∫ ∞f(t)δ ′ (t) dt = f(t)δ(t) − f ′ (t)δ(t) dt−∞ −∞= −f ′ (0), (13.19)<strong>and</strong> similarly <strong>for</strong> higher derivatives.For many practical purposes, effects that are not strictly described by a δ-function may be analysed as such, if they take place in an interval much shorterthan the response interval of the system on which they act. For example, theidealised notion of an impulse of magnitude J applied at time t 0 can be representedbyj(t) =Jδ(t − t 0 ). (13.20)Many physical situations are described by a δ-function in space rather than intime. Moreover, we often require the δ-function to be defined in more than onedimension. For example, the charge density of a point charge q at a point r 0 maybe expressed as a three-dimensional δ-functionρ(r) =qδ(r − r 0 )=qδ(x − x 0 )δ(y − y 0 )δ(z − z 0 ), (13.21)so that a discrete ‘quantum’ is expressed as if it were a continuous distribution.From (13.21) we see that (as expected) the total charge enclosed in a volume Vis given by∫∫{q if r0 lies in V,ρ(r) dV = qδ(r − r 0 ) dV =VV0 otherwise.Closely related to the Dirac δ-function is the Heaviside or unit step functionH(t), <strong>for</strong> which{1 <strong>for</strong> t>0,H(t) =0 <strong>for</strong> t


INTEGRAL TRANSFORMS◮Prove relation (13.23).Considering the integral∫ ∞−∞[ ] ∞ ∫ ∞f(t)H ′ (t) dt = f(t)H(t) − f ′ (t)H(t) dt−∞ −∞∫ ∞= f(∞) − f ′ (t) dt= f(∞) −0[f(t) = f(0),<strong>and</strong> comparing it with (13.12) when a = 0 immediately shows that H ′ (t) =δ(t). ◭] ∞013.1.4 Relation of the δ-function to Fourier trans<strong>for</strong>msIn the previous section we introduced the Dirac δ-function as a way of representingvery sharp narrow pulses, but in no way related it to Fourier trans<strong>for</strong>ms.We now show that the δ-function can equally well be defined in a way that morenaturally relates it to the Fourier trans<strong>for</strong>m.Referring back to the Fourier inversion theorem (13.4), we havef(t) = 1 ∫ ∞ ∫ ∞dω e iωt du f(u) e −iωu=2π−∞∫ ∞−∞du f(u){ 12π−∞∫ ∞−∞}e iω(t−u) dω .Comparison of this with (13.12) shows that we may write the δ-function asδ(t − u) = 1 ∫ ∞e iω(t−u) dω. (13.24)2π −∞Considered as a Fourier trans<strong>for</strong>m, this representation shows that a verynarrow time peak at t = u results from the superposition of a complete spectrumof harmonic waves, all frequencies having the same amplitude <strong>and</strong> all waves beingin phase at t = u. This suggests that the δ-function may also be represented asthe limit of the trans<strong>for</strong>m of a uni<strong>for</strong>m distribution of unit height as the widthof this distribution becomes infinite.Consider the rectangular distribution of frequencies shown in figure 13.4(a).From (13.6), taking the inverse Fourier trans<strong>for</strong>m,f Ω (t) = √ 1 ∫ Ω1 × e iωt dω2π −Ω= √ 2Ω sin Ωt. (13.25)2π ΩtThis function is illustrated in figure 13.4(b) <strong>and</strong> it is apparent that, <strong>for</strong> large Ω, itbecomes very large at t = 0 <strong>and</strong> also very narrow about t = 0, as we qualitatively442


13.1 FOURIER TRANSFORMS˜fΩ2Ω(2π) 1/2f Ω (t)1−Ω(a)ΩωπΩ(b)tFigure 13.4 (a) A Fourier trans<strong>for</strong>m showing a rectangular distribution offrequencies between ±Ω; (b) the function of which it is the trans<strong>for</strong>m, whichis proportional to t −1 sin Ωt.expect <strong>and</strong> require. We also note that, in the limit Ω →∞, f Ω (t), as defined bythe inverse Fourier trans<strong>for</strong>m, tends to (2π) 1/2 δ(t) by virtue of (13.24). Hence wemay conclude that the δ-function can also be represented byδ(t) = limΩ→∞( sin Ωtπt). (13.26)Several other function representations are equally valid, e.g. the limiting cases ofrectangular, triangular or Gaussian distributions; the only essential requirementsare a knowledge of the area under such a curve <strong>and</strong> that undefined operationssuch as dividing by zero are not inadvertently carried out on the δ-function whilstsome non-explicit representation is being employed.We also note that the Fourier trans<strong>for</strong>m definition of the delta function, (13.24),shows that the latter is real sinceδ ∗ (t) = 1 ∫ ∞e −iωt dω = δ(−t) =δ(t).2π −∞Finally, the Fourier trans<strong>for</strong>m of a δ-function is simply˜δ(ω) = 1 √2π∫ ∞−∞δ(t) e −iωt dt = 1 √2π. (13.27)13.1.5 Properties of Fourier trans<strong>for</strong>msHaving considered the Dirac δ-function, we now return to our discussion of theproperties of Fourier trans<strong>for</strong>ms. As we would expect, Fourier trans<strong>for</strong>ms havemany properties analogous to those of Fourier series in respect of the connectionbetween the trans<strong>for</strong>ms of related functions. Here we list these properties withoutproof; they can be verified by working from the definition of the trans<strong>for</strong>m. Aspreviously, we denote the Fourier trans<strong>for</strong>m of f(t) by˜f(ω) orF[f(t)].443


INTEGRAL TRANSFORMS(i) Differentiation:F [ f ′ (t) ] = iω˜f(ω). (13.28)This may be extended to higher derivatives, so that<strong>and</strong>soon.(ii) Integration:F [ f ′′ (t) ] = iωF [ f ′ (t) ] = −ω 2˜f(ω),t]F[∫f(s) ds = 1iω ˜f(ω)+2πcδ(ω), (13.29)where the term 2πcδ(ω) represents the Fourier trans<strong>for</strong>m of the constantof integration associated with the indefinite integral.(iii) Scaling:F[f(at)] = 1 ( ω). (13.30)a˜fa(iv) Translation:(v) Exponential multiplication:where α may be real, imaginary or complex.F[f(t + a)] = e iaω˜f(ω). (13.31)F [ e αt f(t) ] = ˜f(ω + iα), (13.32)◮Prove relation (13.28).Calculating the Fourier trans<strong>for</strong>m of f ′ (t) directly, we obtainF [ f ′ (t) ] = √ 1 ∫ ∞f ′ (t) e −iωt dt2π −∞= √ 1 [e −iωt f(t) + 1 ∫ ∞√ iω e −iωt f(t) dt2π 2π] ∞−∞= iω˜f(ω),if f(t) → 0att = ±∞, asitmustsince ∫ ∞|f(t)| dt is finite. ◭−∞To illustrate a use <strong>and</strong> also a proof of (13.32), let us consider an amplitudemodulatedradio wave. Suppose a message to be broadcast is represented by f(t).The message can be added electronically to a constant signal a of magnitudesuch that a + f(t) is never negative, <strong>and</strong> then the sum can be used to modulatethe amplitude of a carrier signal of frequency ω c . Using a complex exponentialnotation, the transmitted amplitude is now−∞g(t) =A [a + f(t)] e iωct . (13.33)444


13.1 FOURIER TRANSFORMSIgnoring in the present context the effect of the term Aa exp(iω c t), which gives acontribution to the transmitted spectrum only at ω = ω c , we obtain <strong>for</strong> the newspectrum˜g(ω) = √ 1 ∫ ∞A f(t) e iωct e −iωt dt2π −∞= √ 1 ∫ ∞A f(t) e −i(ω−ωc)t dt2π−∞= A˜f(ω − ω c ), (13.34)which is simply a shift of the whole spectrum by the carrier frequency. The useof different carrier frequencies enables signals to be separated.13.1.6 Odd <strong>and</strong> even functionsIf f(t) is odd or even then we may derive alternative <strong>for</strong>ms of Fourier’s inversiontheorem, which lead to the definition of different trans<strong>for</strong>m pairs. Let us firstconsider an odd function f(t) =−f(−t), whose Fourier trans<strong>for</strong>m is given by˜f(ω) = √ 1 ∫ ∞f(t) e −iωt dt2π −∞= √ 1 ∫ ∞f(t)(cos ωt − i sin ωt) dt2π −∞= √ −2i ∫ ∞f(t)sinωt dt,2π0where in the last line we use the fact that f(t) <strong>and</strong>sinωt are odd, whereas cos ωtis even.We note that ˜f(−ω) =−˜f(ω), i.e. ˜f(ω) is an odd function of ω. Hencef(t) = √ 1 ∫ ∞˜f(ω) e iωt dω =2i ∫ ∞√ ˜f(ω)sinωt dω2π −∞2π 0= 2 ∫ ∞{∫ ∞}dω sin ωt f(u)sinωu du .π 00Thus we may define the Fourier sine trans<strong>for</strong>m pair <strong>for</strong> odd functions:√ ∫ 2 ∞˜f s (ω) = f(t) sinωt dt, (13.35)π 0√ ∫ 2 ∞f(t) = ˜fs (ω) sinωt dω. (13.36)π 0Note that although the Fourier sine trans<strong>for</strong>m pair was derived by consideringan odd function f(t) defined over all t, the definitions (13.35) <strong>and</strong> (13.36) onlyrequire f(t) <strong>and</strong>˜f s (ω) to be defined <strong>for</strong> positive t <strong>and</strong> ω respectively. For an445


INTEGRAL TRANSFORMSg(y)(a)(b)(c)(d)0yFigure 13.5 Resolution functions: (a) ideal δ-function; (b) typical unbiasedresolution; (c) <strong>and</strong> (d) biases tending to shift observations to higher valuesthan the true one.even function, i.e. one <strong>for</strong> which f(t) =f(−t), we can define the Fourier cosinetrans<strong>for</strong>m pair in a similar way, but with sin ωt replaced by cos ωt.13.1.7 Convolution <strong>and</strong> deconvolutionIt is apparent that any attempt to measure the value of a physical quantity islimited, to some extent, by the finite resolution of the measuring apparatus used.On the one h<strong>and</strong>, the physical quantity we wish to measure will be in general afunction of an independent variable, x say, i.e. the true function to be measuredtakes the <strong>for</strong>m f(x). On the other h<strong>and</strong>, the apparatus we are using does not givethe true output value of the function; a resolution function g(y) is involved. Bythis we mean that the probability that an output value y = 0 will be recordedinstead as being between y <strong>and</strong> y+dy is given by g(y) dy. Some possible resolutionfunctions of this sort are shown in figure 13.5. To obtain good results we wishthe resolution function to be as close to a δ-function as possible (case (a)). Atypical piece of apparatus has a resolution function of finite width, although ifit is accurate the mean is centred on the true value (case (b)). However, someapparatus may show a bias that tends to shift observations to higher or lowervalues than the true ones (cases (c) <strong>and</strong>(d)), thereby exhibiting systematic error.Given that the true distribution is f(x) <strong>and</strong> the resolution function of ourmeasuring apparatus is g(y), we wish to calculate what the observed distributionh(z) will be. The symbols x, y <strong>and</strong> z all refer to the same physical variable (e.g.446


13.1 FOURIER TRANSFORMSf(x)∗ g(y) =12bh(z)2b−aaFigure 13.6x y z−b b−aaThe convolution of two functions f(x) <strong>and</strong>g(y).length or angle), but are denoted differently because the variable appears in theanalysis in three different roles.The probability that a true reading lying between x <strong>and</strong> x + dx, <strong>and</strong> so havingprobability f(x) dx of being selected by the experiment, will be moved by theinstrumental resolution by an amount z − x into a small interval of width dz isg(z − x) dz. Hence the combined probability that the interval dx will give rise toan observation appearing in the interval dz is f(x) dx g(z − x) dz. Adding togetherthe contributions from all values of x thatcanleadtoanobservationintherangez to z + dz, we find that the observed distribution is given byh(z) =∫ ∞−∞f(x)g(z − x) dx. (13.37)The integral in (13.37) is called the convolution of the functions f <strong>and</strong> g <strong>and</strong> isoften written f ∗ g. The convolution defined above is commutative (f ∗ g = g ∗ f),associative <strong>and</strong> distributive. The observed distribution is thus the convolution ofthe true distribution <strong>and</strong> the experimental resolution function. The result will bethat the observed distribution is broader <strong>and</strong> smoother than the true one <strong>and</strong>, ifg(y) has a bias, the maxima will normally be displaced from their true positions.It is also obvious from (13.37) that if the resolution is the ideal δ-function,g(y) =δ(y) thenh(z) =f(z) <strong>and</strong> the observed distribution is the true one.It is interesting to note, <strong>and</strong> a very important property, that the convolution ofany function g(y) with a number of delta functions leaves a copy of g(y) attheposition of each of the delta functions.◮Find the convolution of the function f(x) =δ(x + a) +δ(x − a) with the function g(y)plotted in figure 13.6.Using the convolution integral (13.37)h(z) =∫ ∞−∞f(x)g(z − x) dx =∫ ∞−∞[δ(x + a)+δ(x − a)]g(z − x) dx= g(z + a)+g(z − a).This convolution h(z) is plotted in figure 13.6. ◭Let us now consider the Fourier trans<strong>for</strong>m of the convolution (13.37); this is447


INTEGRAL TRANSFORMSgiven by∫1 ∞{∫ ∞}˜h(k) = √ dz e −ikz f(x)g(z − x) dx2π −∞−∞= √ 1 ∫ ∞{∫ ∞}dx f(x) g(z − x) e −ikz dz .2π −∞−∞If we let u = z − x in the second integral we have∫1 ∞{∫ ∞}˜h(k) = √ dx f(x) g(u) e −ik(u+x) du2π −∞−∞= √ 1 ∫ ∞∫ ∞f(x) e −ikx dx g(u) e −iku du2π−∞−∞= 1 √2π× √ 2π ˜f(k) × √ 2π˜g(k) = √ 2π ˜f(k)˜g(k). (13.38)Hence the Fourier trans<strong>for</strong>m of a convolution f ∗ g is equal to the product of theseparate Fourier trans<strong>for</strong>ms multiplied by √ 2π; this result is called the convolutiontheorem.It may be proved similarly that the converse is also true, namely that theFourier trans<strong>for</strong>m of the product f(x)g(x) is given byF[f(x)g(x)] = √ 1 ˜f(k) ∗ ˜g(k). (13.39)2π◮Find the Fourier trans<strong>for</strong>m of the function in figure 13.3 representing two wide slits byconsidering the Fourier trans<strong>for</strong>ms of (i) two δ-functions, at x = ±a, (ii) a rectangularfunction of height 1 <strong>and</strong> width 2b centred on x =0.(i) The Fourier trans<strong>for</strong>m of the two δ-functions is given by∫1 ∞˜f(q) = √ δ(x − a) e −iqx dx + 1 ∫ ∞√ δ(x + a) e −iqx dx2π −∞2π −∞= √ 1 (e −iqa + e iqa) = 2cosqa √ .2π 2π(ii) The Fourier trans<strong>for</strong>m of the broad slit is˜g(q) = √ 1 ∫ be −iqx dx = 1 [ ] e−iqx b√2π −b2π −iq−b= −1iq √ 2π (e−iqb − e iqb )= 2sinqbq √ 2π .We have already seen that the convolution of these functions is the required functionrepresenting two wide slits (see figure 13.6). So, using the convolution theorem, the Fouriertrans<strong>for</strong>m of the convolution is √ 2π times the product of the individual trans<strong>for</strong>ms, i.e.4cosqa sin qb/(q √ 2π). This is, of course, the same result as that obtained in the examplein subsection 13.1.2. ◭448


13.1 FOURIER TRANSFORMSThe inverse of convolution, called deconvolution, allows us to find a truedistribution f(x) given an observed distribution h(z) <strong>and</strong> a resolution functiong(y).◮An experimental quantity f(x) is measured using apparatus with a known resolution functiong(y) to give an observed distribution h(z). Howmayf(x) be extracted from the measureddistribution?From the convolution theorem (13.38), the Fourier trans<strong>for</strong>m of the measured distributionis˜h(k) =√2π ˜f(k)˜g(k),from which we obtain1 ˜h(k)˜f(k) = √2π ˜g(k) .Then on inverse Fourier trans<strong>for</strong>ming we find[f(x) = √ 1 ˜h(k)F −1 .2π ˜g(k)In words, to extract the true distribution, we divide the Fourier trans<strong>for</strong>m of the observeddistribution by that of the resolution function <strong>for</strong> each value of k <strong>and</strong> then take the inverseFourier trans<strong>for</strong>m of the function so generated. ◭This explicit method of extracting true distributions is straight<strong>for</strong>ward <strong>for</strong> exactfunctions but, in practice, because of experimental <strong>and</strong> statistical uncertainties inthe experimental data or because data over only a limited range are available, itis often not very precise, involving as it does three (numerical) trans<strong>for</strong>ms eachrequiring in principle an integral over an infinite range.]13.1.8 Correlation functions <strong>and</strong> energy spectraThe cross-correlation of two functions f <strong>and</strong> g is defined byC(z) =∫ ∞−∞f ∗ (x)g(x + z) dx. (13.40)Despite the <strong>for</strong>mal similarity between (13.40) <strong>and</strong> the definition of the convolutionin (13.37), the use <strong>and</strong> interpretation of the cross-correlation <strong>and</strong> of the convolutionare very different; the cross-correlation provides a quantitative measure ofthe similarity of two functions f <strong>and</strong> g as one is displaced through a distance zrelative to the other. The cross-correlation is often notated as C = f ⊗ g, <strong>and</strong>, likeconvolution, it is both associative <strong>and</strong> distributive. Unlike convolution, however,it is not commutative, in fact[f ⊗ g](z) =[g ⊗ f] ∗ (−z). (13.41)449


INTEGRAL TRANSFORMS◮Prove the Wiener–Kinchin theorem,˜C(k) = √ 2π [˜f(k)] ∗˜g(k). (13.42)Following a method similar to that <strong>for</strong> the convolution of f <strong>and</strong> g, let us consider theFourier trans<strong>for</strong>m of (13.40):˜C(k) = √ 1 ∫ ∞{∫ ∞}dz e −ikz f ∗ (x)g(x + z) dx2π −∞−∞= √ 1 ∫ ∞{∫ ∞}dx f ∗ (x) g(x + z) e −ikz dz .2π −∞−∞Making the substitution u = x + z inthesecondintegralweobtain˜C(k) = √ 1 ∫ ∞{∫ ∞}dx f ∗ (x) g(u) e −ik(u−x) du2π −∞−∞= √ 1 ∫ ∞∫ ∞f ∗ (x) e ikx dx g(u) e −iku du2π −∞−∞= √ 1 × √ 2π [˜f(k)] ∗ × √ 2π ˜g(k) = √ 2π [˜f(k)] ∗˜g(k). ◭2πThus the Fourier trans<strong>for</strong>m of the cross-correlation of f <strong>and</strong> g is equal tothe product of [˜f(k)] ∗ <strong>and</strong> ˜g(k) multiplied by √ 2π. This a statement of theWiener–Kinchin theorem. Similarly we can derive the converse theoremF [ f ∗ (x)g(x) ] = √ 1 ˜f ⊗ ˜g.2πIf we now consider the special case where g is taken to be equal to f in (13.40)then, writing the LHS as a(z), we havea(z) =∫ ∞−∞f ∗ (x)f(x + z) dx; (13.43)this is called the auto-correlation function of f(x). Using the Wiener–Kinchintheorem (13.42) we see thata(z) = √ 1 ∫ ∞ã(k) e ikz dk2π −∞= √ 1 ∫ ∞ √2π [˜f(k)]∗˜f(k) e ikz dk,2π−∞so that a(z) is the inverse Fourier trans<strong>for</strong>m of √ 2π |˜f(k)| 2 , which is in turn calledthe energy spectrum of f.13.1.9 Parseval’s theoremUsing the results of the previous section we can immediately obtain Parseval’stheorem. The most general <strong>for</strong>m of this (also called the multiplication theorem) is450


13.1 FOURIER TRANSFORMSobtained simply by noting from (13.42) that the cross-correlation (13.40) of twofunctions f <strong>and</strong> g canbewrittenasC(z) =∫ ∞−∞f ∗ (x)g(x + z) dx =∫ ∞−∞[˜f(k)] ∗˜g(k) e ikz dk. (13.44)Then, setting z = 0 gives the multiplication theorem∫ ∞∫f ∗ (x)g(x) dx = [˜f(k)] ∗˜g(k) dk. (13.45)−∞Specialising further, by letting g = f, we derive the most common <strong>for</strong>m ofParseval’s theorem,∫ ∞∫ ∞|f(x)| 2 dx = |˜f(k)| 2 dk. (13.46)−∞−∞When f is a physical amplitude these integrals relate to the total intensity involvedin some physical process. We have already met a <strong>for</strong>m of Parseval’s theorem <strong>for</strong>Fourier series in chapter 12; it is in fact a special case of (13.46).◮The displacement of a damped harmonic oscillator as a function of time is given by{0 <strong>for</strong> t


INTEGRAL TRANSFORMStwo- or three-dimensional functions of position. For example, in three dimensionswe can define the Fourier trans<strong>for</strong>m of f(x, y, z) as∫∫∫1˜f(k x ,k y ,k z )=f(x, y, z) e −ikxx e −ikyy e −ikzz dx dy dz, (13.47)(2π) 3/2<strong>and</strong> its inverse as∫∫∫1f(x, y, z) =˜f(k(2π) 3/2 x ,k y ,k z ) e ikxx e ikyy e ikzz dk x dk y dk z . (13.48)Denoting the vector with components k x ,k y ,k z by k <strong>and</strong> that with componentsx, y, z by r, we can write the Fourier trans<strong>for</strong>m pair (13.47), (13.48) as∫1˜f(k) = f(r) e −ik·r d 3 r, (13.49)(2π) 3/2 ∫1f(r) = ˜f(k) e ik·r d 3 k. (13.50)(2π) 3/2From these relations we may deduce that the three-dimensional Dirac δ-functioncanbewrittenasδ(r) = 1 ∫(2π) 3 e ik·r d 3 k. (13.51)Similar relations to (13.49), (13.50) <strong>and</strong> (13.51) exist <strong>for</strong> spaces of other dimensionalities.◮In three-dimensional space a function f(r) possesses spherical symmetry, so that f(r) =f(r). Find the Fourier trans<strong>for</strong>m of f(r) as a one-dimensional integral.Let us choose spherical polar coordinates in which the vector k of the Fourier trans<strong>for</strong>mlies along the polar axis (θ =0).Thiswec<strong>and</strong>osincef(r) is spherically symmetric. Wethen haved 3 r = r 2 sin θdrdθdφ <strong>and</strong> k · r = kr cos θ,where k = |k|. The Fourier trans<strong>for</strong>m is then given by∫1˜f(k) = f(r) e −ik·r d 3 r(2π) 3/2 ∫1 ∞ ∫ π ∫ 2π=dr dθ dφ f(r)r 2 −ikr cos θsin θe(2π) 3/2 0 0 0∫1 ∞ ∫ π=dr 2πf(r)r 2 dθ sin θe −ikr cos θ .(2π) 3/2 00The integral over θ may be straight<strong>for</strong>wardly evaluated by noting thatddθ (e−ikr cos θ )=ikr sin θe −ikr cos θ .There<strong>for</strong>e∫1 ∞[ ] e˜f(k) = dr 2πf(r)r 2 −ikr cos θ θ=π(2π) 3/2 0ikrθ=0∫1 ∞( ) sin kr=4πr 2 f(r) dr. ◭(2π) 3/2 kr0452


13.2 LAPLACE TRANSFORMSA similar result may be obtained <strong>for</strong> two-dimensional Fourier trans<strong>for</strong>ms inwhich f(r) =f(ρ), i.e. f(r) is independent of azimuthal angle φ. In this case, usingthe integral representation of the Bessel function J 0 (x) given at the very end ofsubsection 18.5.3, we find˜f(k) = 1 ∫ ∞2πρf(ρ)J 0 (kρ) dρ. (13.52)2π013.2 Laplace trans<strong>for</strong>msOften we are interested in functions f(t) <strong>for</strong> which the Fourier trans<strong>for</strong>m does notexist because f ↛ 0ast →∞, <strong>and</strong> so the integral defining ˜f does not converge.This would be the case <strong>for</strong> the function f(t) =t, which does not possess a Fouriertrans<strong>for</strong>m. Furthermore, we might be interested in a given function only <strong>for</strong> t>0,<strong>for</strong> example when we are given the value at t = 0 in an initial-value problem.This leads us to consider the Laplace trans<strong>for</strong>m, ¯f(s) orL [f(t)], off(t), whichis defined by¯f(s) ≡∫ ∞0f(t)e −st dt, (13.53)provided that the integral exists. We assume here that s is real, but complex valueswould have to be considered in a more detailed study. In practice, <strong>for</strong> a givenfunction f(t) there will be some real number s 0 such that the integral in (13.53)exists <strong>for</strong> s>s 0 but diverges <strong>for</strong> s ≤ s 0 .Through (13.53) we define a linear trans<strong>for</strong>mation L that converts functionsof the variable t to functions of a new variable s:L [af 1 (t)+bf 2 (t)] = aL [f 1 (t)] + bL [f 2 (t)] = a ¯f 1 (s)+b ¯f 2 (s). (13.54)◮Find the Laplace trans<strong>for</strong>ms of the functions (i) f(t) =1, (ii) f(t) =e at , (iii) f(t) =t n ,<strong>for</strong> n =0, 1, 2,... .(i) By direct application of the definition of a Laplace trans<strong>for</strong>m (13.53), we find∫ ∞[ ] ∞ −1L [1] = e −st dt =0s e−st = 1 , if s>0,0swhere the restriction s>0 is required <strong>for</strong> the integral to exist.(ii) Again using (13.53) directly, we find∫ ∞∫ ∞¯f(s) = e at e −st dt = e (a−s)t dt00[ ] e(a−s)t ∞= = 1 if s>a.a − s0s − a453


INTEGRAL TRANSFORMS(iii) Once again using the definition (13.53) we have∫ ∞¯f n (s) = t n e −st dt.0Integrating by parts we find[ ] −t¯f n e −st ∞n (s) =+ n ∫ ∞t n−1 e −st dts0s 0=0+ n s ¯f n−1 (s), if s>0.We now have a recursion relation between successive trans<strong>for</strong>ms <strong>and</strong> by calculating¯f 0 we can infer ¯f 1 , ¯f2 ,etc.Sincet 0 = 1, (i) above gives¯f 0 = 1 , if s>0, (13.55)s<strong>and</strong>¯f 1 (s) = 1 s , 2 ¯f2 (s) = 2!s , ..., 3 ¯fn (s) = n! if s>0.s n+1Thus, in each case (i)–(iii), direct application of the definition of the Laplace trans<strong>for</strong>m(13.53) yields the required result. ◭Unlike that <strong>for</strong> the Fourier trans<strong>for</strong>m, the inversion of the Laplace trans<strong>for</strong>mis not an easy operation to per<strong>for</strong>m, since an explicit <strong>for</strong>mula <strong>for</strong> f(t), given ¯f(s),is not straight<strong>for</strong>wardly obtained from (13.53). The general method <strong>for</strong> obtainingan inverse Laplace trans<strong>for</strong>m makes use of complex variable theory <strong>and</strong> is notdiscussed until chapter 25. However, progress can be made without having to findan explicit inverse, since we can prepare from (13.53) a ‘dictionary’ of the Laplacetrans<strong>for</strong>ms of common functions <strong>and</strong>, when faced with an inversion to carry out,hope to find the given trans<strong>for</strong>m (together with its parent function) in the listing.Such a list is given in table 13.1.When finding inverse Laplace trans<strong>for</strong>ms using table 13.1, it is useful to notethat <strong>for</strong> all practical purposes the inverse Laplace trans<strong>for</strong>m is unique § <strong>and</strong> linearso thatL −1[ a ¯f 1 (s)+b ¯f 2 (s) ] = af 1 (t)+bf 2 (t). (13.56)In many practical problems the method of partial fractions can be useful inproducing an expression from which the inverse Laplace trans<strong>for</strong>m can be found.◮Using table 13.1 find f(t) if¯f(s) = s +3s(s +1) .Using partial fractions ¯f(s) may be written¯f(s) = 3 s − 2s +1 .§ This is not strictly true, since two functions can differ from one another at a finite number ofisolated points but have the same Laplace trans<strong>for</strong>m.454


13.2 LAPLACE TRANSFORMSf(t) ¯f(s) s0c c/s 0ct n cn!/s n+1 0sin bt b/(s 2 + b 2 ) 0cos bt s/(s 2 + b 2 ) 0e at 1/(s − a) at n e at n!/(s − a) n+1 asinh at a/(s 2 − a 2 ) |a|cosh at s/(s 2 − a 2 ) |a|e at sin bt b/[(s − a) 2 + b 2 ] ae at cos bt (s − a)/[(s − a) 2 + b 2 ] at 1/2 1 2 (π/s3 ) 1/2 0t −1/2 (π/s) 1/2 0δ(t − t 0 ) e −st 0 0{1 <strong>for</strong> t ≥ t 0H(t − t 0 )=e −st 0/s 00 <strong>for</strong> ts 0 .Comparing this with the st<strong>and</strong>ard Laplace trans<strong>for</strong>ms in table 13.1, we find that the inversetrans<strong>for</strong>m of 3/s is 3 <strong>for</strong> s>0 <strong>and</strong> the inverse trans<strong>for</strong>m of 2/(s +1)is 2e −t <strong>for</strong> s>−1,<strong>and</strong> sof(t) =3− 2e −t , if s>0. ◭13.2.1 Laplace trans<strong>for</strong>ms of derivatives <strong>and</strong> integralsOne of the main uses of Laplace trans<strong>for</strong>ms is in solving differential equations.Differential equations are the subject of the next six chapters <strong>and</strong> we will returnto the application of Laplace trans<strong>for</strong>ms to their solution in chapter 15. Inthe meantime we will derive the required results, i.e. the Laplace trans<strong>for</strong>ms ofderivatives.The Laplace trans<strong>for</strong>m of the first derivative of f(t) is given byL[ dfdt]=∫ ∞0dfdt e−st dt= [ f(t)e −st] ∫ ∞∞0 + s f(t)e −st dt0= −f(0) + s ¯f(s), <strong>for</strong> s>0. (13.57)The evaluation relies on integration by parts <strong>and</strong> higher-order derivatives maybe found in a similar manner.455


INTEGRAL TRANSFORMS◮Find the Laplace trans<strong>for</strong>m of d 2 f/dt 2 .Using the definition of the Laplace trans<strong>for</strong>m <strong>and</strong> integrating by parts we obtain[ ] dL2 ∫f ∞d 2 f=dt 2 dt 2 e−st dt0[ ] ∞ ∫ df∞=dt e−st + s0 0= − dfdt (0) + s[s ¯f(s) − f(0)],dfdt e−st dt<strong>for</strong> s>0,where (13.57) has been substituted <strong>for</strong> the integral. This can be written more neatly as[ ] dL2 f= s 2 df¯f(s) − sf(0) −dt 2 dt (0), <strong>for</strong> s>0. ◭In general the Laplace trans<strong>for</strong>m of the nth derivative is given by[ d n ]fLdt n = s n ¯f − s n−1 n−2 dff(0) − sdt (0) − ···− dn−1 f(0), <strong>for</strong> s>0.dtn−1 (13.58)We now turn to integration, which is much more straight<strong>for</strong>ward. From thedefinition (13.53),[∫ t] ∫ ∞ ∫ tL f(u) du = dt e −st f(u) du000=[− 1 ∫ t] ∞ ∫ ∞1s e−st f(u) du +00 0 s e−st f(t) dt.The first term on the RHS vanishes at both limits, <strong>and</strong> so[∫ t]L f(u) du = 1 L [f] . (13.59)s013.2.2 Other properties of Laplace trans<strong>for</strong>msFrom table 13.1 it will be apparent that multiplying a function f(t) bye at has theeffect on its trans<strong>for</strong>m that s is replaced by s − a. This is easily proved generally:L [ e at f(t) ] ==∫ ∞0∫ ∞0f(t)e at e −st dtf(t)e −(s−a)t dt= ¯f(s − a). (13.60)As it were, multiplying f(t) bye at moves the origin of s by an amount a.456


13.2 LAPLACE TRANSFORMSWe may now consider the effect of multiplying the Laplace trans<strong>for</strong>m ¯f(s) bye −bs (b>0). From the definition (13.53),∫ ∞e −bs ¯f(s) = e −s(t+b) f(t) dt=0∫ ∞0e −sz f(z − b) dz,on putting t + b = z. Thus e −bs ¯f(s) is the Laplace trans<strong>for</strong>m of a function g(t)defined by{0 <strong>for</strong> 0 b.In other words, the function f has been translated to ‘later’ t (larger values of t)by an amount b.Further properties of Laplace trans<strong>for</strong>ms can be proved in similar ways <strong>and</strong>are listed below.(i)L [f(at)] = 1 a ¯f( s, (13.61)a)(ii)(iii)L [t n n dn ¯f(s)f(t)] =(−1)ds n , <strong>for</strong> n =1, 2, 3,..., (13.62)[ ] ∫ f(t) ∞L = ¯f(u) du, (13.63)t sprovided lim t→0 [f(t)/t] exists.Related results may be easily proved.◮Find an expression <strong>for</strong> the Laplace trans<strong>for</strong>m of td 2 f/dt 2 .From the definition of the Laplace trans<strong>for</strong>m we have[ ] ∫L t d2 f ∞= e −st t d2 fdt 2 dt dt 20∫ ∞= − d e −st d2 fds 0 dt dt 2= − d ds [s2 ¯f(s) − sf(0) − f ′ (0)]= −s 2 d ¯fds − 2s ¯f + f(0). ◭Finally we mention the convolution theorem <strong>for</strong> Laplace trans<strong>for</strong>ms (which isanalogous to that <strong>for</strong> Fourier trans<strong>for</strong>ms discussed in subsection 13.1.7). If thefunctions f <strong>and</strong> g have Laplace trans<strong>for</strong>ms ¯f(s) <strong>and</strong>ḡ(s) then[∫ t]L f(u)g(t − u) du = ¯f(s)ḡ(s), (13.64)0457


INTEGRAL TRANSFORMSFigure 13.7text).Two representations of the Laplace trans<strong>for</strong>m convolution (seewhere the integral in the brackets on the LHS is the convolution of f <strong>and</strong> g,denoted by f ∗ g. As in the case of Fourier trans<strong>for</strong>ms, the convolution definedabove is commutative, i.e. f ∗ g = g ∗ f, <strong>and</strong> is associative <strong>and</strong> distributive. From(13.64) we also see thatL −1[ ∫] t¯f(s)ḡ(s) = f(u)g(t − u) du = f ∗ g.0◮Prove the convolution theorem (13.64) <strong>for</strong> Laplace trans<strong>for</strong>ms.From the definition (13.64),¯f(s)ḡ(s) ==∫ ∞0∫ ∞0e −su f(u) dudu∫ ∞0∫ ∞0e −sv g(v) dvdv e −s(u+v) f(u)g(v).Now letting u + v = t changes the limits on the integrals, with the result that¯f(s)ḡ(s) =∫ ∞0du f(u)∫ ∞udt g(t − u) e −st .As shown in figure 13.7(a) the shaded area of integration may be considered as the sumof vertical strips. However, we may instead integrate over this area by summing overhorizontal strips as shown in figure 13.7(b). Then the integral can be written as¯f(s)ḡ(s) ==∫ t0∫ ∞0∫ ∞du f(u) dt g(t − u) e −st0{∫ t}dt e −st f(u)g(t − u) du[∫ t]= L f(u)g(t − u) du . ◭04580


13.3 CONCLUDING REMARKSThe properties of the Laplace trans<strong>for</strong>m derived in this section can sometimesbe useful in finding the Laplace trans<strong>for</strong>ms of particular functions.◮Find the Laplace trans<strong>for</strong>m of f(t) =t sin bt.Although we could calculate the Laplace trans<strong>for</strong>m directly, we can use (13.62) to give¯f(s) =(−1) d ds L [sin bt] = − d ( ) b 2bs= , <strong>for</strong> s>0. ◭ds s 2 + b 2 (s 2 + b 2 )213.3 Concluding remarksIn this chapter we have discussed Fourier <strong>and</strong> Laplace trans<strong>for</strong>ms in some detail.Both are examples of integral trans<strong>for</strong>ms, which can be considered in a moregeneral context.A general integral trans<strong>for</strong>m of a function f(t) takes the <strong>for</strong>mF(α) =∫ baK(α, t)f(t) dt, (13.65)where F(α) is the trans<strong>for</strong>m of f(t) with respect to the kernel K(α, t), <strong>and</strong> α isthe trans<strong>for</strong>m variable. For example, in the Laplace trans<strong>for</strong>m case K(s, t) =e −st ,a =0,b = ∞.Very often the inverse trans<strong>for</strong>m can also be written straight<strong>for</strong>wardly <strong>and</strong>we obtain a trans<strong>for</strong>m pair similar to that encountered in Fourier trans<strong>for</strong>ms.Examples of such pairs are(i) the Hankel trans<strong>for</strong>mF(k) =f(x) =∫ ∞0∫ ∞0f(x)J n (kx)xdx,F(k)J n (kx)kdk,where the J n are Bessel functions of order n, <strong>and</strong>(ii) the Mellin trans<strong>for</strong>mF(z) =∫ ∞0f(t) = 12πit z−1 f(t) dt,∫ i∞−i∞t −z F(z) dz.Although we do not have the space to discuss their general properties, thereader should at least be aware of this wider class of integral trans<strong>for</strong>ms.459


INTEGRAL TRANSFORMS13.4 Exercises13.1 Find the Fourier trans<strong>for</strong>m of the function f(t) =exp(−|t|).(a) By applying Fourier’s inversion theorem prove that∫π∞2 exp(−|t|) = cos ωt0 1+ω dω. 2(b) By making the substitution ω =tanθ, demonstrate the validity of Parseval’stheorem <strong>for</strong> this function.13.2 Use the general definition <strong>and</strong> properties of Fourier trans<strong>for</strong>ms to show thefollowing.(a) If f(x) is periodic with period a then ˜f(k) = 0, unless ka =2πn <strong>for</strong> integer n.(b) The Fourier trans<strong>for</strong>m of tf(t) isid˜f(ω)/dω.(c) The Fourier trans<strong>for</strong>m of f(mt + c) ise iωc/mm ˜f( ω).m13.3 Find the Fourier trans<strong>for</strong>m of H(x−a)e −bx ,whereH(x) is the Heaviside function.13.4 Prove that the Fourier trans<strong>for</strong>m of the function f(t) defined in the tf-plane bystraight-line segments joining (−T,0) to (0, 1) to (T,0), with f(t) = 0 outside|t|


13.4 EXERCISESDetermine the convolution of f with itself <strong>and</strong>, without further integration,deduce its trans<strong>for</strong>m. Deduce that∫ ∞sin 2 ωdω = π,ω 2−∞∫ ∞−∞sin 4 ωω 4 dω = 2π 3 .13.8 Calculate the Fraunhofer spectrum produced by a diffraction grating, uni<strong>for</strong>mlyilluminated by light of wavelength 2π/k, as follows. Consider a grating with 4Nequal strips each of width a <strong>and</strong> alternately opaque <strong>and</strong> transparent. The aperturefunction is then{A <strong>for</strong> (2n +1)a ≤ y ≤ (2n +2)a, −N ≤ n


INTEGRAL TRANSFORMS13.10 In many applications in which the frequency spectrum of an analogue signal isrequired, the best that can be done is to sample the signal f(t) a finite number oftimes at fixed intervals, <strong>and</strong> then use a discrete Fourier trans<strong>for</strong>m F k to estimatediscrete points on the (true) frequency spectrum ˜f(ω).(a) By an argument that is essentially the converse of that given in section 13.1,show that, if N samples f n , beginning at t = 0 <strong>and</strong> spaced τ apart, are taken,then ˜f(2πk/(Nτ)) ≈ F k τ whereF k = √ 1 ∑N−1f n e −2πnki/N .2π(b) For the function f(t) defined by{1 <strong>for</strong> 0 ≤ t


13.4 EXERCISES(a) Find the Fourier trans<strong>for</strong>m off(γ, p, t) ={e −γt sin pt t > 0,0 t 0) <strong>and</strong> p are constant parameters.(b) The current I(t) flowing through a certain system is related to the appliedvoltage V (t) by the equationI(t) =∫ ∞−∞K(t − u)V (u) du,whereK(τ) =a 1 f(γ 1 ,p 1 ,τ)+a 2 f(γ 2 ,p 2 ,τ).The function f(γ, p, t) is as given in (a) <strong>and</strong> all the a i ,γ i (> 0) <strong>and</strong> p i are fixedparameters. By considering the Fourier trans<strong>for</strong>m of I(t), find the relationshipthat must hold between a 1 <strong>and</strong> a 2 if the total net charge Q passed throughthe system (over a very long time) is to be zero <strong>for</strong> an arbitrary appliedvoltage.13.14 Prove the equality∫ ∞e −2at sin 2 at dt = 1 ∫ ∞a 20π 0 4a 4 + ω dω. 413.15 A linear amplifier produces an output that is the convolution of its input <strong>and</strong> itsresponse function. The Fourier trans<strong>for</strong>m of the response function <strong>for</strong> a particularamplifier isiω˜K(ω) = √ .2π(α + iω)2Determine the time variation of its output g(t) when its input is the Heavisidestep function. (Consider the Fourier trans<strong>for</strong>m of a decaying exponential function<strong>and</strong> the result of exercise 13.2(b).)13.16 In quantum mechanics, two equal-mass particles having momenta p j = k j <strong>and</strong>energies E j = ω j <strong>and</strong> represented by plane wavefunctions φ j =exp[i(k j·r j −ω j t)],j =1, 2, interact through a potential V = V (|r 1 − r 2 |). In first-order perturbationtheory the probability of scattering to a state with momenta <strong>and</strong> energies p ′ j ,E′ jis determined by the modulus squared of the quantity∫∫∫M = ψf ∗ Vψ i dr 1 dr 2 dt.The initial state, ψ i ,isφ 1 φ 2 <strong>and</strong> the final state, ψ f ,isφ ′ 1 φ′ 2 .(a) By writing r 1 + r 2 =2R <strong>and</strong> r 1 − r 2 = r <strong>and</strong> assuming that dr 1 dr 2 = dR dr,show that M can be written as the product of three one-dimensional integrals.(b) From two of the integrals deduce energy <strong>and</strong> momentum conservation in the<strong>for</strong>m of δ-functions.(c) Show that M is proportional to the Fourier trans<strong>for</strong>m of V ,i.e.toṼ (k)where 2k =(p 2 − p 1 ) − (p ′ 2 − p′ 1 ) or, alternatively, k = p′ 1 − p 1.13.17 For some ion–atom scattering processes, the potential V of the previous exercisemay be approximated by V = |r 1 − r 2 | −1 exp(−µ|r 1 − r 2 |). Show, using the resultof the worked example in subsection 13.1.10, that the probability that the ionwill scatter from, say, p 1 to p ′ 1 is proportional to (µ2 + k 2 ) −2 ,wherek = |k| <strong>and</strong> kis as given in part (c) of that exercise.463


INTEGRAL TRANSFORMS13.18 The equivalent duration <strong>and</strong> b<strong>and</strong>width, T e <strong>and</strong> B e , of a signal x(t) are definedin terms of the latter <strong>and</strong> its Fourier trans<strong>for</strong>m ˜x(ω) byT e = 1 ∫ ∞x(t) dt,x(0) −∞B e = 1 ∫ ∞˜x(ω) dω,˜x(0) −∞where neither x(0) nor ˜x(0) is zero. Show that the product T e B e =2π (this is a<strong>for</strong>m of uncertainty principle), <strong>and</strong> find the equivalent b<strong>and</strong>width of the signalx(t) =exp(−|t|/T ).For this signal, determine the fraction of the total energy that lies in the frequencyrange |ω| |a|;2464


13.4 EXERCISES(c) L [sinh at cos bt] = a(s 2 − a 2 + b 2 )[(s − a) 2 + b 2 ] −1 [(s + a) 2 + b 2 ] −1 .13.24 Find the solution (the so-called impulse response or Green’s function) of theequationT dxdt + x = δ(t)by proceeding as follows.(a) Show by substitution thatx(t) =A(1 − e −t/T )H(t)is a solution, <strong>for</strong> which x(0) = 0, ofT dxdt + x = AH(t), (∗)where H(t) is the Heaviside step function.(b) Construct the solution when the RHS of (∗) is replaced by AH(t − τ), withdx/dt = x =0<strong>for</strong>t0, the function y(t) obeys the differential equationd 2 ydt + a dy2 dt + by = c cos2 ωt,where a, b <strong>and</strong> c are positive constants. Find ȳ(s) <strong>and</strong>showthatsȳ(s) → c/2bas s → 0. Interpret the result in the t-domain.13.26 By writing f(x) as an integral involving the δ-function δ(ξ − x) <strong>and</strong> taking theLaplace trans<strong>for</strong>ms of both sides, show that the trans<strong>for</strong>m of the solution of theequationd 4 ydx − y = f(x)4<strong>for</strong> which y <strong>and</strong> its first three derivatives vanish at x = 0 can be written as∫ ∞ȳ(s) = f(ξ) e−sξ0 s 4 − 1 dξ.Use the properties of Laplace trans<strong>for</strong>ms <strong>and</strong> the entries in table 13.1 to showthaty(x) = 1 ∫ xf(ξ) [sinh(x − ξ) − sin(x − ξ)] dξ.20465


INTEGRAL TRANSFORMS13.27 The function f a (x) is defined as unity <strong>for</strong> 0


13.5 HINTS AND ANSWERS13.17 Ṽ (k) ∝ [−2π/(ik)] ∫ {exp[−(µ − ik)r] − exp[−(µ + ik)r]} dr.13.19 Note that the lower limit in the calculation of a(z) is0,<strong>for</strong>z>0, <strong>and</strong> |z|, <strong>for</strong>z


14First-order ordinary differentialequationsDifferential equations are the group of equations that contain derivatives. Chapters14–21 discuss a variety of differential equations, starting in this chapter <strong>and</strong>the next with those ordinary differential equations (ODEs) that have closed-<strong>for</strong>msolutions. As its name suggests, an ODE contains only ordinary derivatives (nopartial derivatives) <strong>and</strong> describes the relationship between these derivatives ofthe dependent variable, usually called y, with respect to the independent variable,usually called x. The solution to such an ODE is there<strong>for</strong>e a function of x <strong>and</strong>is written y(x). For an ODE to have a closed-<strong>for</strong>m solution, it must be possibleto express y(x) in terms of the st<strong>and</strong>ard elementary functions such as exp x, lnx,sin x etc. The solutions of some differential equations cannot, however, be writtenin closed <strong>for</strong>m, but only as an infinite series; these are discussed in chapter 16.Ordinary differential equations may be separated conveniently into differentcategories according to their general characteristics. The primary groupingadopted here is by the order of the equation. The order of an ODE is simply theorder of the highest derivative it contains. Thus equations containing dy/dx, butno higher derivatives, are called first order, those containing d 2 y/dx 2 are calledsecond order <strong>and</strong> so on. In this chapter we consider first-order equations, <strong>and</strong> inthe next, second- <strong>and</strong> higher-order equations.Ordinary differential equations may be classified further according to degree.The degree of an ODE is the power to which the highest-order derivative israised, after the equation has been rationalised to contain only integer powers ofderivatives. Hence the ODEd 3 ( )y dy 3/2dx 3 + x + x 2 y =0,dxis of third order <strong>and</strong> second degree, since after rationalisation it contains the term(d 3 y/dx 3 ) 2 .The general solution to an ODE is the most general function y(x) that satisfiesthe equation; it will contain constants of integration which may be determined by468


14.1 GENERAL FORM OF SOLUTIONthe application of some suitable boundary conditions. For example, we may betold that <strong>for</strong> a certain first-order differential equation, the solution y(x) isequaltozero when the parameter x is equal to unity; this allows us to determine the valueof the constant of integration. The general solutions to nth-order ODEs, whichare considered in detail in the next chapter, will contain n (essential) arbitraryconstants of integration <strong>and</strong> there<strong>for</strong>e we will need n boundary conditions if theseconstants are to be determined (see section 14.1). When the boundary conditionshave been applied, <strong>and</strong> the constants found, we are left with a particular solutionto the ODE, which obeys the given boundary conditions. Some ODEs of degreegreater than unity also possess singular solutions, which are solutions that containno arbitrary constants <strong>and</strong> cannot be found from the general solution; singularsolutions are discussed in more detail in section 14.3. When any solution to anODE has been found, it is always possible to check its validity by substitutioninto the original equation <strong>and</strong> verification that any given boundary conditionsare met.In this chapter, firstly we discuss various types of first-degree ODE <strong>and</strong> then goon to examine those higher-degree equations that can be solved in closed <strong>for</strong>m.At the outset, however, we discuss the general <strong>for</strong>m of the solutions of ODEs;this discussion is relevant to both first- <strong>and</strong> higher-order ODEs.14.1 General <strong>for</strong>m of solutionIt is helpful when considering the general <strong>for</strong>m of the solution of an ODE toconsider the inverse process, namely that of obtaining an ODE from a givengroup of functions, each one of which is a solution of the ODE. Suppose themembers of the group can be written asy = f(x, a 1 ,a 2 ,...,a n ), (14.1)each member being specified by a different set of values of the parameters a i .Forexample, consider the group of functionsy = a 1 sin x + a 2 cos x; (14.2)here n =2.Since an ODE is required <strong>for</strong> which any of the group is a solution, it clearlymust not contain any of the a i . As there are n of the a i in expression (14.1), wemust obtain n + 1 equations involving them in order that, by elimination, we canobtain one final equation without them.Initially we have only (14.1), but if this is differentiated n times, a total of n +1equations is obtained from which (in principle) all the a i can be eliminated, togive one ODE satisfied by all the group. As a result of the n differentiations,d n y/dx n will be present in one of the n + 1 equations <strong>and</strong> hence in the finalequation, which will there<strong>for</strong>e be of nth order.469


FIRST-ORDER ORDINARY DIFFERENTIAL EQUATIONSIn the case of (14.2), we havedydx = a 1 cos x − a 2 sin x,d 2 ydx 2 = −a 1 sin x − a 2 cos x.Here the elimination of a 1 <strong>and</strong> a 2 is trivial (because of the similarity of the <strong>for</strong>msof y <strong>and</strong> d 2 y/dx 2 ), resulting ind 2 ydx 2 + y =0,a second-order equation.Thus, to summarise, a group of functions (14.1) with n parameters satisfies annth-order ODE in general (although in some degenerate cases an ODE of lessthan nth order is obtained). The intuitive converse of this is that the generalsolution of an nth-order ODE contains n arbitrary parameters (constants); <strong>for</strong>our purposes, this will be assumed to be valid although a totally general proof isdifficult.As mentioned earlier, external factors affect a system described by an ODE,by fixing the values of the dependent variables <strong>for</strong> particular values of theindependent ones. These externally imposed (or boundary) conditions on thesolution are thus the means of determining the parameters <strong>and</strong> so of specifyingprecisely which function is the required solution. It is apparent that the numberof boundary conditions should match the number of parameters <strong>and</strong> hence theorder of the equation, if a unique solution is to be obtained. Fewer independentboundary conditions than this will lead to a number of undetermined parametersin the solution, whilst an excess will usually mean that no acceptable solution ispossible.For an nth-order equation the required n boundary conditions can take many<strong>for</strong>ms, <strong>for</strong> example the value of y at n different values of x, or the value of anyn − 1ofthen derivatives dy/dx, d 2 y/dx 2 , ..., d n y/dx n together with that of y, all<strong>for</strong> the same value of x, or many intermediate combinations.14.2 First-degree first-order equationsFirst-degree first-order ODEs contain only dy/dx equated to some function of x<strong>and</strong> y, <strong>and</strong> can be written in either of two equivalent st<strong>and</strong>ard <strong>for</strong>ms,dy= F(x, y), A(x, y) dx + B(x, y) dy =0,dxwhere F(x, y) =−A(x, y)/B(x, y), <strong>and</strong> F(x, y), A(x, y) <strong>and</strong>B(x, y) are in generalfunctions of both x <strong>and</strong> y. Which of the two above <strong>for</strong>ms is the more useful<strong>for</strong> finding a solution depends on the type of equation being considered. There470


14.2 FIRST-DEGREE FIRST-ORDER EQUATIONSare several different types of first-degree first-order ODEs that are of interest inthe physical sciences. These equations <strong>and</strong> their respective solutions are discussedbelow.14.2.1 Separable-variable equationsA separable-variable equation is one which may be written in the conventional<strong>for</strong>mdy= f(x)g(y), (14.3)dxwhere f(x) <strong>and</strong>g(y) are functions of x <strong>and</strong> y respectively, including cases inwhich f(x) org(y) is simply a constant. Rearranging this equation so that theterms depending on x <strong>and</strong> on y appear on opposite sides (i.e. are separated), <strong>and</strong>integrating, we obtain∫ ∫ dyg(y) = f(x) dx.Finding the solution y(x) that satisfies (14.3) then depends only on the ease withwhich the integrals in the above equation can be evaluated. It is also worthnoting that ODEs that at first sight do not appear to be of the <strong>for</strong>m (14.3) cansometimes be made separable by an appropriate factorisation.◮Solvedydx = x + xy.Since the RHS of this equation can be factorised to give x(1 + y), the equation becomesseparable <strong>and</strong> we obtain∫ ∫dy1+y = xdx.Now integrating both sides separately, we findln(1 + y) = x22 + c,<strong>and</strong> so( ) x21+y =exp2 + c = A expwhere c <strong>and</strong> hence A is an arbitrary constant. ◭Solution method. Factorise the equation so that it becomes separable. Rearrangeit so that the terms depending on x <strong>and</strong> those depending on y appear on oppositesides <strong>and</strong> then integrate directly. Remember the constant of integration, which canbe evaluated if further in<strong>for</strong>mation is given.( x22),471


FIRST-ORDER ORDINARY DIFFERENTIAL EQUATIONS14.2.2 Exact equationsAn exact first-degree first-order ODE is one of the <strong>for</strong>mA(x, y) dx + B(x, y) dy = 0 <strong>and</strong> <strong>for</strong> which ∂A∂y = ∂B∂x . (14.4)In this case A(x, y) dx + B(x, y) dy is an exact differential, dU(x, y) say (seesection 5.3). In other wordsAdx+ Bdy= dU = ∂U ∂Udx +∂x ∂y dy,from which we obtainA(x, y) = ∂U∂x , (14.5)B(x, y) = ∂U∂y . (14.6)Since ∂ 2 U/∂x∂y = ∂ 2 U/∂y∂x we there<strong>for</strong>e require∂A∂y = ∂B∂x . (14.7)If (14.7) holds then (14.4) can be written dU(x, y) = 0, which has the solutionU(x, y) =c, wherec is a constant <strong>and</strong> from (14.5) U(x, y) is given by∫U(x, y) = A(x, y) dx + F(y). (14.8)The function F(y) can be found from (14.6) by differentiating (14.8) with respectto y <strong>and</strong> equating to B(x, y).◮Solvex dy +3x + y =0.dxRearranging into the <strong>for</strong>m (14.4) we have(3x + y) dx + xdy =0,i.e. A(x, y) =3x + y <strong>and</strong> B(x, y) =x. Since∂A/∂y =1=∂B/∂x, the equation is exact, <strong>and</strong>by (14.8) the solution is given by∫U(x, y) = (3x + y) dx + F(y) =c 1 ⇒ 3x22 + yx + F(y) =c 1.Differentiating U(x, y) with respect to y <strong>and</strong> equating it to B(x, y) =x we obtain dF/dy =0,which integrates immediately to give F(y) =c 2 . There<strong>for</strong>e, letting c = c 1 − c 2 ,thesolutionto the original ODE is3x 2+ xy = c. ◭2472


14.2 FIRST-DEGREE FIRST-ORDER EQUATIONSSolution method. Check that the equation is an exact differential using (14.7) thensolve using (14.8). Find the function F(y) by differentiating (14.8) with respect toy <strong>and</strong> using (14.6).14.2.3 Inexact equations: integrating factorsEquations that may be written in the <strong>for</strong>mA(x, y) dx + B(x, y) dy = 0 but <strong>for</strong> which ∂A∂y ≠ ∂B∂x(14.9)are known as inexact equations. However, the differential Adx+ Bdy can alwaysbe made exact by multiplying by an integrating factor µ(x, y), which obeys∂(µA)∂y= ∂(µB)∂x . (14.10)For an integrating factor that is a function of both x <strong>and</strong> y, i.e.µ = µ(x, y), thereexists no general method <strong>for</strong> finding it; in such cases it may sometimes be foundby inspection. If, however, an integrating factor exists that is a function of eitherx or y alone then (14.10) can be solved to find it. For example, if we assumethat the integrating factor is a function of x alone, i.e. µ = µ(x), then (14.10)readsµ ∂A∂y = µ∂B ∂x + B dµdx .Rearranging this expression we finddµµ = 1 ( ∂AB ∂y − ∂B )dx = f(x) dx,∂xwhere we require f(x) also to be a function of x only; indeed this provides ageneral method of determining whether the integrating factor µ is a function ofx alone. This integrating factor is then given by{∫µ(x) =exp}f(x) dxwheref(x) = 1 B( ∂A∂y − ∂B ). (14.11)∂xSimilarly, if µ = µ(y) then{∫µ(y) =exp}g(y) dywhere473g(y) = 1 A( ∂B∂x − ∂A ). (14.12)∂y


FIRST-ORDER ORDINARY DIFFERENTIAL EQUATIONS◮Solvedydx = − 2 y − 3y2x .Rearranging into the <strong>for</strong>m (14.9), we have(4x +3y 2 ) dx +2xy dy =0, (14.13)i.e. A(x, y) =4x +3y 2 <strong>and</strong> B(x, y) =2xy. Now∂A∂y =6y, ∂B∂x =2y,so the ODE is not exact in its present <strong>for</strong>m. However, we see that(1 ∂AB ∂y − ∂B )= 2 ∂x x ,a function of x alone. There<strong>for</strong>e an integrating factor exists that is also a function of xalone <strong>and</strong>, ignoring the arbitrary constant of integration, is given by{ ∫ } dxµ(x) =exp 2 =exp(2lnx) =x 2 .xMultiplying (14.13) through by µ(x) =x 2 we obtain(4x 3 +3x 2 y 2 ) dx +2x 3 ydy=4x 3 dx +(3x 2 y 2 dx +2x 3 ydy)=0.By inspection this integrates immediately to give the solution x 4 + y 2 x 3 = c, wherec is aconstant. ◭Solution method. Examine whether f(x) <strong>and</strong> g(y) are functions of only x or yrespectively. If so, then the required integrating factor is a function of either x ory only, <strong>and</strong> is given by (14.11) or (14.12) respectively. If the integrating factor isa function of both x <strong>and</strong> y, then sometimes it may be found by inspection or bytrial <strong>and</strong> error. In any case, the integrating factor µ must satisfy (14.10). Once theequation has been made exact, solve by the method of subsection 14.2.2.14.2.4 Linear equationsLinear first-order ODEs are a special case of inexact ODEs (discussed in theprevious subsection) <strong>and</strong> can be written in the conventional <strong>for</strong>mdy+ P (x)y = Q(x). (14.14)dxSuch equations can be made exact by multiplying through by an appropriateintegrating factor in a similar manner to that discussed above. In this case,however, the integrating factor is always a function of x alone <strong>and</strong> may beexpressed in a particularly simple <strong>for</strong>m. An integrating factor µ(x) must be suchthatµ(x) dyd+ µ(x)P (x)y = [µ(x)y] = µ(x)Q(x), (14.15)dx dx474


14.2 FIRST-DEGREE FIRST-ORDER EQUATIONSwhich may then be integrated directly to give∫µ(x)y = µ(x)Q(x) dx. (14.16)The required integrating factor µ(x) is determined by the first equality in (14.15),i.e.d dy(µy) =µdx dx + dµdx y = µ dydx + µPy,which immediately gives the simple relation{∫dµdx = µ(x)P (x) ⇒ µ(x) =exp}P (x) dx . (14.17)◮Solvedy+2xy =4x.dxThe integrating factor is given immediately by{∫ }µ(x) =exp 2xdx =expx 2 .Multiplying through the ODE by µ(x) =expx 2 <strong>and</strong> integrating, we have∫y exp x 2 =4 x exp x 2 dx =2expx 2 + c.The solution to the ODE is there<strong>for</strong>e given by y =2+c exp(−x 2 ). ◭Solution method. Rearrange the equation into the <strong>for</strong>m (14.14) <strong>and</strong> multiply by theintegrating factor µ(x) given by (14.17). The left- <strong>and</strong> right-h<strong>and</strong> sides can then beintegrated directly, giving y from (14.16).14.2.5 Homogeneous equationsHomogeneous equation are ODEs that may be written in the <strong>for</strong>mdy A(x, y)( y)=dx B(x, y) = F , (14.18)xwhere A(x, y) <strong>and</strong> B(x, y) are homogeneous functions of the same degree. Afunction f(x, y) is homogeneous of degree n if, <strong>for</strong> any λ, itobeysf(λx, λy) =λ n f(x, y).For example, if A = x 2 y − xy 2 <strong>and</strong> B = x 3 + y 3 then we see that A <strong>and</strong> B areboth homogeneous functions of degree 3. In general, <strong>for</strong> functions of the <strong>for</strong>m ofA <strong>and</strong> B, we see that <strong>for</strong> both to be homogeneous, <strong>and</strong> of the same degree, werequire the sum of the powers in x <strong>and</strong> y in each term of A <strong>and</strong> B to be the same475


FIRST-ORDER ORDINARY DIFFERENTIAL EQUATIONS(in this example equal to 3). The RHS of a homogeneous ODE can be written asa function of y/x. The equation may then be solved by making the substitutiony = vx, sothatdydx = v + x dvdx = F(v).This is now a separable equation <strong>and</strong> can be integrated directly to give∫∫dv dxF(v) − v = x . (14.19)◮Solvedydx = y ( y)x +tan .xSubstituting y = vx we obtainv + x dvdx = v +tanv.Cancelling v on both sides, rearranging <strong>and</strong> integrating gives∫∫ dxcot vdv=x =lnx + c 1.But∫∫cot vdv=cos vsin v dv =ln(sinv)+c 2,so the solution to the ODE is y = x sin −1 Ax, whereA is a constant. ◭Solution method. Check to see whether the equation is homogeneous. If so, makethe substitution y = vx, separate variables as in (14.19) <strong>and</strong> then integrate directly.Finally replace v by y/x to obtain the solution.14.2.6 Isobaric equationsAn isobaric ODE is a generalisation of the homogeneous ODE discussed in theprevious section, <strong>and</strong> is of the <strong>for</strong>mdy A(x, y)=dx B(x, y) , (14.20)where the equation is dimensionally consistent if y <strong>and</strong> dy are each given a weightm relative to x <strong>and</strong> dx, i.e. if the substitution y = vx m makes it separable.476


14.2 FIRST-DEGREE FIRST-ORDER EQUATIONS◮Solvedydx = −1 (y 2 + 2 ).2yx xRearranging we have(y 2 + 2 )dx +2yx dy =0.xGiving y <strong>and</strong> dy the weight m <strong>and</strong> x <strong>and</strong> dx the weight 1, the sums of the powers in eachterm on the LHS are 2m +1, 0 <strong>and</strong>2m + 1 respectively. These are equal if 2m +1=0, i.e.if m = − 1 2 . Substituting y = vxm = vx −1/2 , with the result that dy = x −1/2 dv − 1 2 vx−3/2 dx,we obtainvdv+ dx x =0,which is separable <strong>and</strong> may be integrated directly to give 1 2 v2 +lnx = c. Replacingv byy √ x we obtain the solution 1 2 y2 x +lnx = c. ◭Solution method. Write the equation in the <strong>for</strong>m Adx+ Bdy =0. Giving y <strong>and</strong>dy each a weight m <strong>and</strong> x <strong>and</strong> dx each a weight 1, write down the sum of powersin each term. Then, if a value of m that makes all these sums equal can be found,substitute y = vx m into the original equation to make it separable. Integrate theseparated equation directly, <strong>and</strong> then replace v by yx −m to obtain the solution.Bernoulli’s equation has the <strong>for</strong>m14.2.7 Bernoulli’s equationdydx + P (x)y = Q(x)yn where n ≠0or1. (14.21)This equation is very similar in <strong>for</strong>m to the linear equation (14.14), but is in factnon-linear due to the extra y n factor on the RHS. However, the equation can bemade linear by substituting v = y 1−n <strong>and</strong> correspondingly( )dy yn dvdx = 1 − n dx .Substituting this into (14.21) <strong>and</strong> dividing through by y n , we finddv+(1− n)P (x)v =(1− n)Q(x),dxwhich is a linear equation <strong>and</strong> may be solved by the method described insubsection 14.2.4.477


FIRST-ORDER ORDINARY DIFFERENTIAL EQUATIONS◮Solvedydx + y x =2x3 y 4 .If we let v = y 1−4 = y −3 thendydx = − y4 dv3 dx .Substituting this into the ODE <strong>and</strong> rearranging, we obtaindvdx − 3vx = −6x3 ,which is linear <strong>and</strong> may be solved by multiplying through by the integrating factor (seesubsection 14.2.4){ ∫exp −3dxx}=exp(−3lnx) = 1 x 3 .This yields the solutionv= −6x + c.x3 Remembering that v = y −3 ,weobtainy −3 = −6x 4 + cx 3 . ◭Solution method. Rearrange the equation into the <strong>for</strong>m (14.21) <strong>and</strong> make the substitutionv = y 1−n . This leads to a linear equation in v, which can be solved by themethod of subsection 14.2.4. Then replace v by y 1−n to obtain the solution.14.2.8 Miscellaneous equationsThere are two further types of first-degree first-order equation that occur fairlyregularly but do not fall into any of the above categories. They may be reducedto one of the above equations, however, by a suitable change of variable.Firstly, we considerdy= F(ax + by + c), (14.22)dxwhere a, b <strong>and</strong> c are constants, i.e. x <strong>and</strong> y only appear on the RHS in the particularcombination ax + by + c <strong>and</strong> not in any other combination or by themselves. Thisequation can be solved by making the substitution v = ax + by + c, inwhichcasedvdx = a + b dy = a + bF(v), (14.23)dxwhich is separable <strong>and</strong> may be integrated directly.478


14.2 FIRST-DEGREE FIRST-ORDER EQUATIONS◮Solvedydx =(x + y +1)2 .Making the substitution v = x + y + 1, we obtain, as in (14.23),dvdx = v2 +1,which is separable <strong>and</strong> integrates directly to give∫ ∫dv1+v = dx ⇒ tan −1 v = x + c 2 1 .So the solution to the original ODE is tan −1 (x + y +1)=x + c 1 ,wherec 1 is a constant ofintegration. ◭Solution method. In an equation such as (14.22), substitute v = ax+by+c to obtaina separable equation that can be integrated directly. Then replace v by ax + by + cto obtain the solution.Secondly, we discussdy ax + by + c=dx ex + fy + g , (14.24)where a, b, c, e, f <strong>and</strong> g are all constants. This equation may be solved by lettingx = X + α <strong>and</strong> y = Y + β, whereα <strong>and</strong> β are constants found fromaα + bβ + c = 0 (14.25)eα + fβ + g =0. (14.26)Then (14.24) can be written asdY aX + bY=dX eX + fY ,which is homogeneous <strong>and</strong> can be solved by the method of subsection 14.2.5.Note, however, that if a/e = b/f then (14.25) <strong>and</strong> (14.26) are not independent<strong>and</strong> so cannot be solved uniquely <strong>for</strong> α <strong>and</strong> β. However, in this case, (14.24)reduces to an equation of the <strong>for</strong>m (14.22), which was discussed above.◮Solvedy 2x − 5y +3=dx 2x +4y − 6 .Let x = X + α <strong>and</strong> y = Y + β, whereα <strong>and</strong> β obey the relations2α − 5β +3=02α +4β − 6=0,which solve to give α = β = 1. Making these substitutions we finddY 2X − 5Y=dX 2X +4Y ,479


FIRST-ORDER ORDINARY DIFFERENTIAL EQUATIONSwhich is a homogeneous ODE <strong>and</strong> can be solved by substituting Y = vX (see subsection14.2.5) to obtaindvdX2 − 7v − 4v2= .X(2 + 4v)This equation is separable, <strong>and</strong> using partial fractions we find∫2+4v2 − 7v − 4v dv = − 4 ∫dv2 3 4v − 1 − 2 ∫ ∫dv dX3 v +2 = X ,which integrates to giveln X + 1 ln(4v − 1) + 2 ln(v +2)=c 3 3 1,orX 3 (4v − 1)(v +2) 2 =exp3c 1 .Remembering that Y = vX, x = X +1 <strong>and</strong>y = Y + 1, the solution to the original ODEis given by (4y − x − 3)(y +2x − 3) 2 = c 2 ,wherec 2 =exp3c 1 . ◭Solution method. If in (14.24) a/e ≠ b/f then make the substitution x = X + α,y = Y + β, whereα <strong>and</strong> β are given by (14.25) <strong>and</strong> (14.26); the resulting equationis homogeneous <strong>and</strong> can be solved as in subsection 14.2.5. Substitute v = Y/X,X = x − α <strong>and</strong> Y = y − β to obtain the solution. If a/e = b/f then (14.24) is ofthe same <strong>for</strong>m as (14.22) <strong>and</strong> may be solved accordingly.14.3 Higher-degree first-order equationsFirst-order equations of degree higher than the first do not occur often inthe description of physical systems, since squared <strong>and</strong> higher powers of firstorderderivatives usually arise from resistive or driving mechanisms, when anacceleration or other higher-order derivative is also present. They do sometimesappear in connection with geometrical problems, however.Higher-degree first-order equations can be written as F(x, y, dy/dx) =0.Themost general st<strong>and</strong>ard <strong>for</strong>m isp n + a n−1 (x, y)p n−1 + ···+ a 1 (x, y)p + a 0 (x, y) =0, (14.27)where <strong>for</strong> ease of notation we write p = dy/dx. If the equation can be solved <strong>for</strong>one of x, y or p then either an explicit or a parametric solution can sometimes beobtained. We discuss the main types of such equations below, including Clairaut’sequation, which is a special case of an equation explicitly soluble <strong>for</strong> y.14.3.1 Equations soluble <strong>for</strong> pSometimes the LHS of (14.27) can be factorised into the <strong>for</strong>m(p − F 1 )(p − F 2 ) ···(p − F n )=0, (14.28)480


14.3 HIGHER-DEGREE FIRST-ORDER EQUATIONSwhere F i = F i (x, y). We are then left with solving the n first-degree equationsp = F i (x, y). Writing the solutions to these first-degree equations as G i (x, y) =0,the general solution to (14.28) is given by the productG 1 (x, y)G 2 (x, y) ···G n (x, y) =0. (14.29)◮Solve(x 3 + x 2 + x +1)p 2 − (3x 2 +2x +1)yp +2xy 2 =0. (14.30)This equation may be factorised to give[(x +1)p − y][(x 2 +1)p − 2xy] =0.Taking each bracket in turn we have(x +1) dydx − y =0,(x 2 +1) dy − 2xy =0,dxwhich have the solutions y − c(x +1) = 0 <strong>and</strong> y − c(x 2 + 1) = 0 respectively (seesection 14.2 on first-degree first-order equations). Note that the arbitrary constants inthese two solutions can be taken to be the same, since only one is required <strong>for</strong> a first-orderequation. The general solution to (14.30) is then given by[y − c(x +1)] [ y − c(x 2 +1) ] =0. ◭Solution method. If the equation can be factorised into the <strong>for</strong>m (14.28) then solvethe first-order ODE p − F i =0<strong>for</strong> each factor <strong>and</strong> write the solution in the <strong>for</strong>mG i (x, y) =0. The solution to the original equation is then given by the product(14.29).14.3.2 Equations soluble <strong>for</strong> xEquations that can be solved <strong>for</strong> x, i.e. such that they may be written in the <strong>for</strong>mx = F(y, p), (14.31)can be reduced to first-degree first-order equations in p by differentiating bothsides with respect to y, sothatdxdy = 1 p = ∂F∂y + ∂F dp∂p dy .This results in an equation of the <strong>for</strong>m G(y, p) = 0, which can be used togetherwith (14.31) to eliminate p <strong>and</strong> give the general solution. Note that often a singularsolution to the equation will be found at the same time (see the introduction tothis chapter).481


FIRST-ORDER ORDINARY DIFFERENTIAL EQUATIONS◮Solve6y 2 p 2 +3xp − y =0. (14.32)This equation can be solved <strong>for</strong> x explicitly to give 3x =(y/p) − 6y 2 p. Differentiating bothsides with respect to y, we find3 dxdy = 3 p = 1 p − y dp dp−p 2 6y2dy dy − 12yp,which factorises to give( ) ( 1+6yp22p + y dp )=0. (14.33)dySetting the factor containing dp/dy equal to zero gives a first-degree first-order equationin p, which may be solved to give py 2 = c. Substituting <strong>for</strong> p in (14.32) then yields thegeneral solution of (14.32):y 3 =3cx +6c 2 . (14.34)If we now consider the first factor in (14.33), we find 6p 2 y = −1 as a possible solution.Substituting <strong>for</strong> p in (14.32) we find the singular solution8y 3 +3x 2 =0.Note that the singular solution contains no arbitrary constants <strong>and</strong> cannot be found fromthe general solution (14.34) by any choice of the constant c. ◭Solution method. Write the equation in the <strong>for</strong>m (14.31) <strong>and</strong> differentiate bothsides with respect to y. Rearrange the resulting equation into the <strong>for</strong>m G(y, p) =0,which can be used together with the original ODE to eliminate p <strong>and</strong> so give thegeneral solution. If G(y, p) can be factorised then the factor containing dp/dy shouldbe used to eliminate p <strong>and</strong> give the general solution. Using the other factors in thisfashion will instead lead to singular solutions.14.3.3 Equations soluble <strong>for</strong> yEquations that can be solved <strong>for</strong> y, i.e. are such that they may be written in the<strong>for</strong>my = F(x, p), (14.35)can be reduced to first-degree first-order equations in p by differentiating bothsides with respect to x, sothatdydx = p = ∂F∂x + ∂F dp∂p dx .This results in an equation of the <strong>for</strong>m G(x, p) = 0, which can be used togetherwith (14.35) to eliminate p <strong>and</strong> give the general solution. An additional (singular)solution to the equation is also often found.482


14.3 HIGHER-DEGREE FIRST-ORDER EQUATIONS◮Solvexp 2 +2xp − y =0. (14.36)This equation can be solved <strong>for</strong> y explicitly to give y = xp 2 +2xp. Differentiating bothsides with respect to x, we finddydxwhich after factorising gives= p =2xpdpdx + p2 +2x dpdx +2p,((p +1) p +2x dp )=0. (14.37)dxTo obtain the general solution of (14.36), we consider the factor containing dp/dx. Thisfirst-degree first-order equation in p has the solution xp 2 = c (see subsection 14.3.1), whichwe then use to eliminate p from (14.36). Thus we find that the general solution to (14.36)is(y − c) 2 =4cx. (14.38)If instead, we set the other factor in (14.37) equal to zero, we obtain the very simplesolution p = −1. Substituting this into (14.36) then givesx + y =0,which is a singular solution to (14.36). ◭Solution method. Write the equation in the <strong>for</strong>m (14.35) <strong>and</strong> differentiate bothsides with respect to x. Rearrange the resulting equation into the <strong>for</strong>m G(x, p) =0,which can be used together with the original ODE to eliminate p <strong>and</strong> so give thegeneral solution. If G(x, p) can be factorised then the factor containing dp/dx shouldbe used to eliminate p <strong>and</strong> give the general solution. Using the other factors in thisfashion will instead lead to singular solutions.14.3.4 Clairaut’s equationFinally, we consider Clairaut’s equation, which has the <strong>for</strong>my = px + F(p) (14.39)<strong>and</strong> is there<strong>for</strong>e a special case of equations soluble <strong>for</strong> y, as in (14.35). It may besolved by a similar method to that given in subsection 14.3.3, but <strong>for</strong> Clairaut’sequation the <strong>for</strong>m of the general solution is particularly simple. Differentiating(14.39) with respect to x, we finddydx = p = p + x dpdx + dF dpdp dx⇒dp ( ) dFdx dp + x =0. (14.40)Considering first the factor containing dp/dx, we finddpdx = d2 ydx 2 =0 ⇒ y = c 1x + c 2 . (14.41)483


FIRST-ORDER ORDINARY DIFFERENTIAL EQUATIONSSince p = dy/dx = c 1 , if we substitute (14.41) into (14.39) we find c 1 x + c 2 =c 1 x + F(c 1 ). There<strong>for</strong>e the constant c 2 is given by F(c 1 ), <strong>and</strong> the general solutionto (14.39) isy = c 1 x + F(c 1 ), (14.42)i.e. the general solution to Clairaut’s equation can be obtained by replacing pin the ODE by the arbitrary constant c 1 . Now, considering the second factor in(14.40), we also havedF+ x =0, (14.43)dpwhich has the <strong>for</strong>m G(x, p) = 0. This relation may be used to eliminate p from(14.39) to give a singular solution.◮Solvey = px + p 2 . (14.44)From (14.42) the general solution is y = cx + c 2 . But from (14.43) we also have 2p + x =0 ⇒ p = −x/2. Substituting this into (14.44) we find the singular solution x 2 +4y =0.◭Solution method. Write the equation in the <strong>for</strong>m (14.39), then the general solutionis given by replacing p by some constant c, as shown in (14.42). Using the relationdF/dp+x =0to eliminate p from the original equation yields the singular solution.14.4 Exercises14.1 A radioactive isotope decays in such a way that the number of atoms present ata given time, N(t), obeys the equationdNdt = −λN.If there are initially N 0 atoms present, find N(t) atlatertimes.14.2 Solve the following equations by separation of the variables:(a) y ′ − xy 3 =0;(b) y ′ tan −1 x − y(1 + x 2 ) −1 =0;(c) x 2 y ′ + xy 2 =4y 2 .14.3 Show that the following equations either are exact or can be made exact, <strong>and</strong>solve them:(a) y(2x 2 y 2 +1)y ′ + x(y 4 +1)=0;(b) 2xy ′ +3x + y =0;(c) (cos 2 x + y sin 2x)y ′ + y 2 =0.14.4 Find the values of α <strong>and</strong> β that make( 1dF(x, y) =x 2 +2 + α )dx +(xy β +1)dyyan exact differential. For these values solve F(x, y) =0.484


14.4 EXERCISES14.5 By finding suitable integrating factors, solve the following equations:(a) (1 − x 2 )y ′ +2xy =(1− x 2 ) 3/2 ;(b) y ′ − y cot x +cosecx =0;(c) (x + y 3 )y ′ = y (treat y as the independent variable).14.6 By finding an appropriate integrating factor, solvedydx = − 2x2 + y 2 + x.xy14.7 Find, in the <strong>for</strong>m of an integral, the solution of the equationα dydt + y = f(t)<strong>for</strong> a general function f(t). Find the specific solutions <strong>for</strong>(a) f(t) =H(t),(b) f(t) =δ(t),(c) f(t) =β −1 e −t/β H(t) withβ


FIRST-ORDER ORDINARY DIFFERENTIAL EQUATIONSFind an expression <strong>for</strong> the velocity v of the mass as a function of time, giventhat it has an initial velocity v 0 .14.13 Using the results about Laplace trans<strong>for</strong>ms given in chapter 13 <strong>for</strong> df/dt <strong>and</strong>tf(t), show, <strong>for</strong> a function y(t) thatsatisfiest dy +(t − 1)y =0dt (∗)with y(0) finite, that ȳ(s) =C(1 + s) −2 <strong>for</strong> some constant C.Given that∞∑y(t) =t + a n t n ,determine C <strong>and</strong> show that a n =(−1) n−1 /(n − 1)!. Compare this result with thatobtained by integrating (∗) directly.14.14 Solvedydx = 1x +2y +1 .14.15 Solvedydx = − x + y3x +3y − 4 .14.16 If u =1+tany, calculated(ln u)/dy; hence find the general solution ofdy=tanx cos y (cos y +siny).dx14.17 Solvex(1 − 2x 2 y) dydx + y =3x2 y 2 ,given that y(1) = 1/2.14.18 A reflecting mirror is made in the shape of the surface of revolution generated byrevolving the curve y(x) about the x-axis. In order that light rays emitted from apoint source at the origin are reflected back parallel to the x-axis, the curve y(x)must obeyyx = 2p1 − p , 2where p = dy/dx. By solving this equation <strong>for</strong> x, find the curve y(x).14.19 Find the curve with the property that at each point on it the sum of the interceptson the x- <strong>and</strong>y-axes of the tangent to the curve (taking account of sign) is equalto 1.14.20 Find a parametric solution of( ) 2 dyx + dydx dx − y =0as follows.(a) Write an equation <strong>for</strong> y in terms of p = dy/dx <strong>and</strong> show thatp = p 2 +(2px +1) dpdx .(b) Using p as the independent variable, arrange this as a linear first-orderequation <strong>for</strong> x.486n=2


14.4 EXERCISES(c) Find an appropriate integrating factor to obtainx = ln p − p + c ,(1 − p) 2which, together with the expression <strong>for</strong> y obtained in (a), gives a parameterisationof the solution.(d) Reverse the roles of x <strong>and</strong> y in steps (a) to (c), putting dx/dy = p −1 ,<strong>and</strong>show that essentially the same parameterisation is obtained.14.21 Using the substitutions u = x 2 <strong>and</strong> v = y 2 , reduce the equation( ) 2 dyxy − (x 2 + y 2 − 1) dy + xy =0dxdxto Clairaut’s <strong>for</strong>m. Hence show that the equation represents a family of conics<strong>and</strong> the four sides of a square.14.22 The action of the control mechanism on a particular system <strong>for</strong> an input f(t) isdescribed, <strong>for</strong> t ≥ 0, by the coupled first-order equations:ẏ +4z = f(t),ż − 2z = ẏ + 1 y. 2Use Laplace trans<strong>for</strong>ms to find the response y(t) of the system to a unit stepinput, f(t) =H(t), given that y(0) = 1 <strong>and</strong> z(0) = 0.Questions 23 to 31 are intended to give the reader practice in choosing an appropriatemethod. The level of difficulty varies within the set; if necessary, the hints maybe consulted <strong>for</strong> an indication of the most appropriate approach.14.23 Find the general solutions of the following:(a)dydx + xya 2 + x = x; dy(b) 2 dx = 4y2x − 2 y2 .14.24 Solve the following first-order equations <strong>for</strong> the boundary conditions given:(a) y ′ − (y/x) =1, y(1) = −1;(b) y ′ − y tan x =1, y(π/4) = 3;(c) y ′ − y 2 /x 2 =1/4, y(1) = 1;(d) y ′ − y 2 /x 2 =1/4, y(1) = 1/2.14.25 An electronic system has two inputs, to each of which a constant unit signal isapplied, but starting at different times. The equations governing the system thustake the <strong>for</strong>mẋ +2y = H(t),ẏ − 2x = H(t − 3).Initially (at t =0),x =1<strong>and</strong>y = 0; find x(t) atlatertimes.14.26 Solve the differential equationsin x dy +2y cos x =1,dxsubject to the boundary condition y(π/2) = 1.14.27 Find the complete solution of( ) 2 dy− y dydx x dx + A x =0,where A is a positive constant.487


FIRST-ORDER ORDINARY DIFFERENTIAL EQUATIONS14.28 Find the solution of(5x + y − 7) dy =3(x + y +1).dx14.29 Find the solution y = y(x) ofx dydx + y − y2=0,x3/2 subject to y(1) = 1.14.30 Find the solution of(2 sin y − x) dydx =tany,if (a) y(0) = 0, <strong>and</strong> (b) y(0) = π/2.14.31 Find the family of solutions ofthat satisfy y(0) = 0.( )d 2 2y dydx + + dy2 dx dx =014.5 Hints <strong>and</strong> answers14.1 N(t) =N 0 exp(−λt).14.3 (a) exact, x 2 y 4 + x 2 + y 2 = c; (b) IF = x −1/2 , x 1/2 (x + y) = c; (c) IF =sec 2 x, y 2 tan x + y = c.14.5 (a) IF = (1 − x 2 ) −2 , y =(1− x 2 )(k +sin −1 x); (b) IF = cosec x, leading toy = k sin x +cosx; (c) exact equation is y −1 (dx/dy) − xy −2 = y, leading tox = y(k + y 2 /2).14.7 y(t) =e ∫ −t/α t α −1 e t′ /α f(t ′ )dt ′ ;(a)y(t) =1− e −t/α ;(b)y(t) =α −1 e −t/α ;(c)y(t) =(e −t/α − e −t/β )/(α − β). It becomes case (b).14.9 Note that, if the angle between the tangent <strong>and</strong> the radius vector is α, thencos α = dr/ds <strong>and</strong> sin α = p/r.14.11 Homogeneous equation, put y = vx to obtain (1 − v)(v 2 +2v +2) −1 dv = x −1 dx;write 1 − v as 2 − (1 + v), <strong>and</strong> v 2 +2v +2as1+(1+v) 2 ;A[x 2 +(x + y) 2 ]=exp { 4tan −1 [(x + y)/x] } .14.13 (1 + s)(dȳ/ds)+2ȳ =0.C = 1; use separation of variables to show directly thaty(t) =te −t .14.15 The equation is of the <strong>for</strong>m of (14.22), set v = x + y; x +3y +2ln(x + y − 2) = A.14.17 The equation is isobaric with weight y = −2; setting y = vx −2 givesv −1 (1 − v) −1 (1 − 2v) dv = x −1 dx; 4xy(1 − x 2 y)=1.14.19 The curve must satisfy y =(1−p −1 ) −1 (1−x+px), which has solution x =(p−1) −2 ,leading to y =(1± √ x) 2 or x =(1± √ y) 2 ; the singular solution p ′ = 0 givesstraight lines joining (θ, 0) <strong>and</strong> (0, 1 − θ) <strong>for</strong>anyθ.14.21 v = qu + q/(q − 1), where q = dv/du. General solution y 2 = cx 2 + c/(c − 1),hyperbolae <strong>for</strong> c>0 <strong>and</strong> ellipses <strong>for</strong> c


14.5 HINTS AND ANSWERS14.31 Show that p =(Ce x − 1) −1 , where p = dy/dx; y =ln[C − e −x )/(C − 1)] orln[D − (D − 1)e −x ]orln(e −K +1− e −x )+K.489


15Higher-order ordinary differentialequationsFollowing on from the discussion of first-order ordinary differential equations(ODEs) given in the previous chapter, we now examine equations of second <strong>and</strong>higher order. Since a brief outline of the general properties of ODEs <strong>and</strong> theirsolutions was given at the beginning of the previous chapter, we will not repeatit here. Instead, we will begin with a discussion of various types of higher-orderequation. This chapter is divided into three main parts. We first discuss linearequations with constant coefficients <strong>and</strong> then investigate linear equations withvariable coefficients. Finally, we discuss a few methods that may be of use insolving general linear or non-linear ODEs. Let us start by considering somegeneral points relating to all linear ODEs.Linear equations are of paramount importance in the description of physicalprocesses. Moreover, it is an empirical fact that, when put into mathematical<strong>for</strong>m, many natural processes appear as higher-order linear ODEs, most oftenas second-order equations. Although we could restrict our attention to thesesecond-order equations, the generalisation to nth-order equations requires littleextra work, <strong>and</strong> so we will consider this more general case.A linear ODE of general order n has the <strong>for</strong>ma n (x) dn ydx n + a n−1(x) dn−1 ydx n−1 + ···+ a 1(x) dydx + a 0(x)y = f(x). (15.1)If f(x) = 0 then the equation is called homogeneous; otherwise it is inhomogeneous.The first-order linear equation studied in subsection 14.2.4 is a special case of(15.1). As discussed at the beginning of the previous chapter, the general solutionto (15.1) will contain n arbitrary constants, which may be determined if n boundaryconditions are also provided.In order to solve any equation of the <strong>for</strong>m (15.1), we must first find thegeneral solution of the complementary equation, i.e. the equation <strong>for</strong>med by setting490


HIGHER-ORDER ORDINARY DIFFERENTIAL EQUATIONSf(x) =0:a n (x) dn ydx n + a n−1(x) dn−1 ydx n−1 + ···+ a 1(x) dydx + a 0(x)y =0. (15.2)To determine the general solution of (15.2), we must find n linearly independentfunctions that satisfy it. Once we have found these solutions, the general solutionis given by a linear superposition of these n functions. In other words, if the nsolutions of (15.2) are y 1 (x), y 2 (x),...,y n (x), then the general solution is given bythe linear superpositiony c (x) =c 1 y 1 (x)+c 2 y 2 (x)+···+ c n y n (x), (15.3)where the c m are arbitrary constants that may be determined if n boundaryconditions are provided. The linear combination y c (x) is called the complementaryfunction of (15.1).The question naturally arises how we establish that any n individual solutions to(15.2) are indeed linearly independent. For n functions to be linearly independentover an interval, there must not exist any set of constants c 1 ,c 2 ,...,c n such thatc 1 y 1 (x)+c 2 y 2 (x)+···+ c n y n (x) = 0 (15.4)over the interval in question, except <strong>for</strong> the trivial case c 1 = c 2 = ···= c n =0.A statement equivalent to (15.4), which is perhaps more useful <strong>for</strong> the practicaldetermination of linear independence, can be found by repeatedly differentiating(15.4), n − 1 times in all, to obtain n simultaneous equations <strong>for</strong> c 1 ,c 2 ,...,c n :c 1 y 1 (x)+c 2 y 2 (x)+···+ c n y n (x) =0c 1 y ′ 1 (x)+c 2 y ′ 2 (x)+···+ c n y ′ n (x) =0.(15.5).c 1 y (n−1)1(x)+c 2 y (n−1)2+ ···+ c n y n (n−1) (x) =0,where the primes denote differentiation with respect to x. Referring to thediscussion of simultaneous linear equations given in chapter 8, if the determinantof the coefficients of c 1 ,c 2 ,...,c n is non-zero then the only solution to equations(15.5) is the trivial solution c 1 = c 2 = ···= c n = 0. In other words, the n functionsy 1 (x), y 2 (x),...,y n (x) are linearly independent over an interval if∣∣W (y 1 ,y 2 ,...,y n )=∣y 1 y 2 ... y n..′y 1′y 2... ....y (n−1)1... ... y n(n−1)≠ 0 (15.6)∣over that interval; W (y 1 ,y 2 ,...,y n ) is called the Wronskian of the set of functions.It should be noted, however, that the vanishing of the Wronskian does notguarantee that the functions are linearly dependent.491


HIGHER-ORDER ORDINARY DIFFERENTIAL EQUATIONSIf the original equation (15.1) has f(x) = 0 (i.e. it is homogeneous) then ofcourse the complementary function y c (x) in (15.3) is already the general solution.If, however, the equation has f(x) ≠ 0 (i.e. it is inhomogeneous) then y c (x) isonlyone part of the solution. The general solution of (15.1) is then given byy(x) =y c (x)+y p (x), (15.7)where y p (x)istheparticular integral,whichcanbeany function that satisfies (15.1)directly, provided it is linearly independent of y c (x). It should be emphasised <strong>for</strong>practical purposes that any such function, no matter how simple (or complicated),is equally valid in <strong>for</strong>ming the general solution (15.7).It is important to realise that the above method <strong>for</strong> finding the general solutionto an ODE by superposing particular solutions assumes crucially that the ODEis linear. For non-linear equations, discussed in section 15.3, this method cannotbe used, <strong>and</strong> indeed it is often impossible to find closed-<strong>for</strong>m solutions to suchequations.15.1 Linear equations with constant coefficientsIf the a m in (15.1) are constants rather than functions of x then we haved n ya ndx n + a d n−1 yn−1dx n−1 + ···+ a dy1dx + a 0y = f(x). (15.8)Equations of this sort are very common throughout the physical sciences <strong>and</strong>engineering, <strong>and</strong> the method <strong>for</strong> their solution falls into two parts as discussedin the previous section, i.e. finding the complementary function y c (x) <strong>and</strong> findingthe particular integral y p (x). If f(x) = 0 in (15.8) then we do not have to finda particular integral, <strong>and</strong> the complementary function is by itself the generalsolution.15.1.1 Finding the complementary function y c (x)The complementary function must satisfyd n ya ndx n + a d n−1 yn−1dx n−1 + ···+ a dy1dx + a 0y = 0 (15.9)<strong>and</strong> contain n arbitrary constants (see equation (15.3)). The st<strong>and</strong>ard method<strong>for</strong> finding y c (x) is to try a solution of the <strong>for</strong>m y = Ae λx , substituting this into(15.9). After dividing the resulting equation through by Ae λx , we are left with apolynomial equation in λ of order n; thisistheauxiliary equation <strong>and</strong> readsa n λ n + a n−1 λ n−1 + ···+ a 1 λ + a 0 =0. (15.10)492


15.1 LINEAR EQUATIONS WITH CONSTANT COEFFICIENTSIn general the auxiliary equation has n roots, say λ 1 ,λ 2 ,...,λ n . In certain cases,some of these roots may be repeated <strong>and</strong> some may be complex. The three maincases are as follows.(i) All roots real <strong>and</strong> distinct. In this case the n solutions to (15.9) are exp λ m x<strong>for</strong> m =1ton. It is easily shown by calculating the Wronskian (15.6)of these functions that if all the λ m are distinct then these solutions arelinearly independent. We can there<strong>for</strong>e linearly superpose them, as in(15.3), to <strong>for</strong>m the complementary functiony c (x) =c 1 e λ1x + c 2 e λ2x + ···+ c n e λnx . (15.11)(ii) Some roots complex. For the special (but usual) case that all the coefficientsa m in (15.9) are real, if one of the roots of the auxiliary equation (15.10)is complex, say α + iβ, then its complex conjugate α − iβ is also a root. Inthis case we can writec 1 e (α+iβ)x + c 2 e (α−iβ)x = e αx (d 1 cos βx + d 2 sin βx)= Ae αx { sincos}(βx + φ), (15.12)where A <strong>and</strong> φ are arbitrary constants.(iii) Some roots repeated. If, <strong>for</strong> example, λ 1 occurs k times (k >1) as a rootof the auxiliary equation, then we have not found n linearly independentsolutions of (15.9); <strong>for</strong>mally the Wronskian (15.6) of these solutions, havingtwo or more identical columns, is equal to zero. We must there<strong>for</strong>e findk −1 further solutions that are linearly independent of those already found<strong>and</strong> also of each other. By direct substitution into (15.9) we find thatxe λ1x , x 2 e λ1x , ..., x k−1 e λ1xare also solutions, <strong>and</strong> by calculating the Wronskian it is easily shown thatthey, together with the solutions already found, <strong>for</strong>m a linearly independentset of n functions. There<strong>for</strong>e the complementary function is given byy c (x) =(c 1 + c 2 x + ···+ c k x k−1 )e λ1x + c k+1 e λk+1x + c k+2 e λk+2x + ···+ c n e λnx .(15.13)If more than one root is repeated the above argument is easily extended.For example, suppose as be<strong>for</strong>e that λ 1 is a k-fold root of the auxiliaryequation <strong>and</strong>, further, that λ 2 is an l-fold root (of course, k>1<strong>and</strong>l>1).Then, from the above argument, the complementary function readsy c (x) =(c 1 + c 2 x + ···+ c k x k−1 )e λ1x+(c k+1 + c k+2 x + ···+ c k+l x l−1 )e λ2x+ c k+l+1 e λk+l+1x + c k+l+2 e λk+l+2x + ···+ c n e λnx . (15.14)493


HIGHER-ORDER ORDINARY DIFFERENTIAL EQUATIONS◮Find the complementary function of the equationd 2 ydx − 2 dy2 dx + y = ex . (15.15)Setting the RHS to zero, substituting y = Ae λx <strong>and</strong> dividing through by Ae λx we obtainthe auxiliary equationλ 2 − 2λ +1=0.The root λ = 1 occurs twice <strong>and</strong> so, although e x is a solution to (15.15), we must finda further solution to the equation that is linearly independent of e x .Fromtheabovediscussion, we deduce that xe x is such a solution, so that the full complementary functionis given by the linear superpositiony c (x) =(c 1 + c 2 x)e x . ◭Solution method. Set the RHS of the ODE to zero (if it is not already so), <strong>and</strong>substitute y = Ae λx . After dividing through the resulting equation by Ae λx , obtainan nth-order polynomial equation in λ (the auxiliary equation, see (15.10)). Solvethe auxiliary equation to find the n roots, λ 1 ,λ 2 ,...,λ n , say. If all these roots arereal <strong>and</strong> distinct then y c (x) is given by (15.11). If, however, some of the roots arecomplex or repeated then y c (x) is given by (15.12) or (15.13), or the extension(15.14) of the latter, respectively.15.1.2 Finding the particular integral y p (x)There is no generally applicable method <strong>for</strong> finding the particular integral y p (x)but, <strong>for</strong> linear ODEs with constant coefficients <strong>and</strong> a simple RHS, y p (x) can oftenbe found by inspection or by assuming a parameterised <strong>for</strong>m similar to f(x). Thelatter method is sometimes called the method of undetermined coefficients. Iff(x)contains only polynomial, exponential, or sine <strong>and</strong> cosine terms then, by assuminga trial function <strong>for</strong> y p (x) of similar <strong>for</strong>m but one which contains a number ofundetermined parameters <strong>and</strong> substituting this trial function into (15.9), theparameters can be found <strong>and</strong> y p (x) deduced. St<strong>and</strong>ard trial functions are asfollows.(i) If f(x) =ae rx then tryy p (x) =be rx .(ii) If f(x) =a 1 sin rx + a 2 cos rx (a 1 or a 2 may be zero) then tryy p (x) =b 1 sin rx + b 2 cos rx.(iii) If f(x) =a 0 + a 1 x + ···+ a N x N (some a m may be zero) then tryy p (x) =b 0 + b 1 x + ···+ b N x N .494


15.1 LINEAR EQUATIONS WITH CONSTANT COEFFICIENTS(iv) If f(x) is the sum or product of any of the above then try y p (x) asthesum or product of the corresponding individual trial functions.It should be noted that this method fails if any term in the assumed trialfunction is also contained within the complementary function y c (x). In such acase the trial function should be multiplied by the smallest integer power of xsuch that it will then contain no term that already appears in the complementaryfunction. The undetermined coefficients in the trial function can now be foundby substitution into (15.8).Three further methods that are useful in finding the particular integral y p (x) arethose based on Green’s functions, the variation of parameters, <strong>and</strong> a change in thedependent variable using knowledge of the complementary function. However,since these methods are also applicable to equations with variable coefficients, adiscussion of them is postponed until section 15.2.◮Find a particular integral of the equationd 2 ydx − 2 dy2 dx + y = ex .From the above discussion our first guess at a trial particular integral would be y p (x) =be x .However, since the complementary function of this equation is y c (x) =(c 1 + c 2 x)e x (asin the previous subsection), we see that e x is already contained in it, as indeed is xe x .Multiplying our first guess by the lowest integer power of x such that the result does notappear in y c (x), we there<strong>for</strong>e try y p (x) =bx 2 e x . Substituting this into the ODE, we findthat b =1/2, so the particular integral is given by y p (x) =x 2 e x /2. ◭Solution method. If the RHS of an ODE contains only functions mentioned at thestart of this subsection then the appropriate trial function should be substitutedinto it, thereby fixing the undetermined parameters. If, however, the RHS of theequation is not of this <strong>for</strong>m then one of the more general methods outlined in subsections15.2.3–15.2.5 should be used; perhaps the most straight<strong>for</strong>ward of these isthe variation-of-parameters method.15.1.3 Constructing the general solution y c (x)+y p (x)As stated earlier, the full solution to the ODE (15.8) is found by adding togetherthe complementary function <strong>and</strong> any particular integral. In order to illustratefurther the material discussed in the last two subsections, let us find the generalsolution to a new example, starting from the beginning.495


HIGHER-ORDER ORDINARY DIFFERENTIAL EQUATIONS◮Solved 2 ydx 2 +4y = x2 sin 2x. (15.16)First we set the RHS to zero <strong>and</strong> assume the trial solution y = Ae λx . Substituting this into(15.16) leads to the auxiliary equationλ 2 +4=0 ⇒ λ = ±2i. (15.17)There<strong>for</strong>e the complementary function is given byy c (x) =c 1 e 2ix + c 2 e −2ix = d 1 cos 2x + d 2 sin 2x. (15.18)We must now turn our attention to the particular integral y p (x). Consulting the list ofst<strong>and</strong>ard trial functions in the previous subsection, we find that a first guess at a suitabletrial function <strong>for</strong> this case should be(ax 2 + bx + c)sin2x +(dx 2 + ex + f)cos2x. (15.19)However, we see that this trial function contains terms in sin 2x <strong>and</strong> cos 2x, both of whichalready appear in the complementary function (15.18). We must there<strong>for</strong>e multiply (15.19)by the smallest integer power of x which ensures that none of the resulting terms appearsin y c (x). Since multiplying by x will suffice, we finally assume the trial function(ax 3 + bx 2 + cx)sin2x +(dx 3 + ex 2 + fx)cos2x. (15.20)Substituting this into (15.16) to fix the constants appearing in (15.20), we find the particularintegral to bey p (x) =− x3 x2cos 2x +12 16 sin 2x + x cos 2x. (15.21)32The general solution to (15.16) then readsy(x) =y c (x)+y p (x)= d 1 cos 2x + d 2 sin 2x − x3 x2cos 2x +12 16 sin 2x + x cos 2x. ◭3215.1.4 Linear recurrence relationsBe<strong>for</strong>e continuing our discussion of higher-order ODEs, we take this opportunityto introduce the discrete analogues of differential equations, which are calledrecurrence relations (or sometimes difference equations). Whereas a differentialequation gives a prescription, in terms of current values, <strong>for</strong> the new value of adependent variable at a point only infinitesimally far away, a recurrence relationdescribes how the next in a sequence of values u n , defined only at (non-negative)integer values of the ‘independent variable’ n, istobecalculated.In its most general <strong>for</strong>m a recurrence relation expresses the way in which u n+1is to be calculated from all the preceding values u 0 ,u 1 ,... ,u n . Just as the mostgeneral differential equations are intractable, so are the most general recurrencerelations, <strong>and</strong> we will limit ourselves to analogues of the types of differentialequations studied earlier in this chapter, namely those that are linear, have496


15.1 LINEAR EQUATIONS WITH CONSTANT COEFFICIENTSconstant coefficients <strong>and</strong> possess simple functions on the RHS. Such equationsoccur over a broad range of engineering <strong>and</strong> statistical physics as well as in therealms of finance, business planning <strong>and</strong> gambling! They <strong>for</strong>m the basis of manynumerical methods, particularly those concerned with the numerical solution o<strong>for</strong>dinary <strong>and</strong> partial differential equations.A general recurrence relation is exemplified by the <strong>for</strong>mulaN−1∑u n+1 = a r u n−r + k, (15.22)r=0where N <strong>and</strong> the a r are fixed <strong>and</strong> k is a constant or a simple function of n.Such an equation, involving terms of the series whose indices differ by up to N(ranging from n−N +1 to n), is called an Nth-order recurrence relation. It is clearthat, given values <strong>for</strong> u 0 ,u 1 ,... ,u N−1 , this is a definitive scheme <strong>for</strong> generating theseries <strong>and</strong> there<strong>for</strong>e has a unique solution.Parallelling the nomenclature of differential equations, if the term not involvingany u n is absent, i.e. k = 0, then the recurrence relation is called homogeneous.The parallel continues with the <strong>for</strong>m of the general solution of (15.22). If v n isthe general solution of the homogeneous relation, <strong>and</strong> w n is any solution of thefull relation, thenu n = v n + w nis the most general solution of the complete recurrence relation. This is straight<strong>for</strong>wardlyverified as follows:u n+1 = v n+1 + w n+1∑N−1=r=0N−1=N−1∑a r v n−r +r=0a r w n−r + k∑a r (v n−r + w n−r )+kr=0∑a r u n−r + k.N−1=r=0Of course, if k =0thenw n = 0 <strong>for</strong> all n is a trivial particular solution <strong>and</strong> thecomplementary solution, v n , is itself the most general solution.First-order recurrence relationsFirst-order relations, <strong>for</strong> which N = 1, are exemplified byu n+1 = au n + k, (15.23)with u 0 specified. The solution to the homogeneous relation is immediate,u n = Ca n ,497


HIGHER-ORDER ORDINARY DIFFERENTIAL EQUATIONS<strong>and</strong>, if k is a constant, the particular solution is equally straight<strong>for</strong>ward: w n = K<strong>for</strong> all n, provided K is chosen to satisfyK = aK + k,i.e. K = k(1 − a) −1 . This will be sufficient unless a =1,inwhichcaseu n = u 0 + nkis obvious by inspection.Thus the general solution of (15.23) is{Ca n + k/(1 − a) a ≠1,u n =u 0 + nk a =1.(15.24)If u 0 is specified <strong>for</strong> the case of a ≠1thenC must be chosen as C = u 0 −k/(1−a),resulting in the equivalent <strong>for</strong>mu n = u 0 a n + k 1 − an1 − a . (15.25)We now illustrate this method with a worked example.◮A house-buyer borrows capital B from a bank that charges a fixed annual rate of interestR%. If the loan is to be repaid over Y years, at what value should the fixed annual paymentsP , made at the end of each year, be set? For a loan over 25 years at 6%, what percentageof the first year’s payment goes towards paying off the capital?Let u n denote the outst<strong>and</strong>ing debt at the end of year n, <strong>and</strong>writeR/100 = r. Then therelevant recurrence relation isu n+1 = u n (1 + r) − Pwith u 0 = B. From (15.25) we haveu n = B(1 + r) n 1 − (1 + r)n− P1 − (1 + r) .AstheloanistoberepaidoverY years, u Y = 0 <strong>and</strong> thusBr(1 + r)YP =(1 + r) Y − 1 .The first year’s interest is rB <strong>and</strong> so the fraction of the first year’s payment goingtowards capital repayment is (P − rB)/P , which, using the above expression <strong>for</strong> P ,isequalto (1 + r) −Y . With the given figures, this is (only) 23%. ◭With only small modifications, the method just described can be adapted toh<strong>and</strong>le recurrence relations in which the constant k in (15.23) is replaced by kα n ,i.e. the relation isu n+1 = au n + kα n . (15.26)As <strong>for</strong> an inhomogeneous linear differential equation (see subsection 15.1.2), wemay try as a potential particular solution a <strong>for</strong>m which resembles the term thatmakes the equation inhomogeneous. Here, the presence of the term kα n indicates498


15.1 LINEAR EQUATIONS WITH CONSTANT COEFFICIENTSthat a particular solution of the <strong>for</strong>m u n = Aα n should be tried. Substituting thisinto (15.26) givesAα n+1 = aAα n + kα n ,from which it follows that A = k/(α − a) <strong>and</strong> that there is a particular solutionhaving the <strong>for</strong>m u n = kα n /(α − a), provided α ≠ a. For the special case α = a, thereader can readily verify that a particular solution of the <strong>for</strong>m u n = Anα n is appropriate.This mirrors the corresponding situation <strong>for</strong> linear differential equationswhen the RHS of the differential equation is contained in the complementaryfunction of its LHS.In summary, the general solution to (15.26) isu n =with C 1 = u 0 − k/(α − a) <strong>and</strong>C 2 = u 0 .{C1 a n + kα n /(α − a) α ≠ a,C 2 a n + knα n−1 α = a,(15.27)Second-order recurrence relationsWe consider next recurrence relations that involve u n−1 in the prescription <strong>for</strong>u n+1 <strong>and</strong> treat the general case in which the intervening term, u n , is also present.A typical equation is thusu n+1 = au n + bu n−1 + k. (15.28)As previously, the general solution of this is u n = v n + w n ,wherev n satisfiesv n+1 = av n + bv n−1 (15.29)<strong>and</strong> w n is any particular solution of (15.28); the proof follows the same lines asthat given earlier.We have already seen <strong>for</strong> a first-order recurrence relation that the solution tothe homogeneous equation is given by terms <strong>for</strong>ming a geometric series, <strong>and</strong> weconsider a corresponding series of powers in the present case. Setting v n = Aλ n in(15.29) <strong>for</strong> some λ, as yet undetermined, gives the requirement that λ should satisfyAλ n+1 = aAλ n + bAλ n−1 .Dividing through by Aλ n−1 (assumed non-zero) shows that λ could be either ofthe roots, λ 1 <strong>and</strong> λ 2 ,ofλ 2 − aλ − b =0, (15.30)which is known as the characteristic equation of the recurrence relation.That there are two possible series of terms of the <strong>for</strong>m Aλ n is consistent with thefact that two initial values (boundary conditions) have to be provided be<strong>for</strong>e theseries can be calculated by repeated use of (15.28). These two values are sufficientto determine the appropriate coefficient A <strong>for</strong> each of the series. Since (15.29) is499


HIGHER-ORDER ORDINARY DIFFERENTIAL EQUATIONSboth linear <strong>and</strong> homogeneous, <strong>and</strong> is satisfied by both v n = Aλ n 1 <strong>and</strong> v n = Bλ n 2 ,itsgeneral solution isv n = Aλ n 1 + Bλ n 2.If the coefficients a <strong>and</strong> b are such that (15.30) has two equal roots, i.e. a 2 = −4b,then, as in the analogous case of repeated roots <strong>for</strong> differential equations (seesubsection 15.1.1(iii)), the second term of the general solution is replaced by Bnλ n 1to givev n =(A + Bn)λ n 1.Finding a particular solution is straight<strong>for</strong>ward if k is a constant: a trivial butadequate solution is w n = k(1 − a − b) −1 <strong>for</strong> all n. As with first-order equations,particular solutions can be found <strong>for</strong> other simple <strong>for</strong>ms of k by trying functionssimilar to k itself. Thus particular solutions <strong>for</strong> the cases k = Cn <strong>and</strong> k = Dα ncan be found by trying w n = E + Fn <strong>and</strong> w n = Gα n respectively.◮Find the value of u 16 if the series u n satisfiesu n+1 +4u n +3u n−1 = n<strong>for</strong> n ≥ 1, withu 0 =1<strong>and</strong> u 1 = −1.We first solve the characteristic equation,λ 2 +4λ +3=0,to obtain the roots λ = −1 <strong>and</strong>λ = −3. Thus the complementary function isv n = A(−1) n + B(−3) n .In view of the <strong>for</strong>m of the RHS of the original relation, we tryas a particular solution <strong>and</strong> obtainyielding F =1/8 <strong>and</strong>E =1/32.Thus the complete general solution isw n = E + FnE + F(n +1)+4(E + Fn)+3[E + F(n − 1)] = n,u n = A(−1) n + B(−3) n + n 8 + 132 ,<strong>and</strong> now using the given values <strong>for</strong> u 0 <strong>and</strong> u 1 determines A as 7/8 <strong>and</strong>B as 3/32. Thusu n = 1 32 [28(−1)n +3(−3) n +4n +1] .Finally, substituting n =16givesu 16 = 4 035 633, a value the reader may (or may not)wish to verify by repeated application of the initial recurrence relation. ◭500


15.1 LINEAR EQUATIONS WITH CONSTANT COEFFICIENTSHigher-order recurrence relationsIt will be apparent that linear recurrence relations of order N>2 do not presentany additional difficulty in principle, though two obvious practical difficulties are(i) that the characteristic equation is of order N <strong>and</strong> in general will not have rootsthat can be written in closed <strong>for</strong>m <strong>and</strong> (ii) that a correspondingly large numberof given values is required to determine the N otherwise arbitrary constants inthe solution. The algebraic labour needed to solve the set of simultaneous linearequations that determines them increases rapidly with N. We do not give specificexamples here, but some are included in the exercises at the end of the chapter.15.1.5 Laplace trans<strong>for</strong>m methodHaving briefly discussed recurrence relations, we now return to the main topicof this chapter, i.e. methods <strong>for</strong> obtaining solutions to higher-order ODEs. Onesuch method is that of Laplace trans<strong>for</strong>ms, which is very useful <strong>for</strong> solvinglinear ODEs with constant coefficients. Taking the Laplace trans<strong>for</strong>m of such anequation trans<strong>for</strong>ms it into a purely algebraic equation in terms of the Laplacetrans<strong>for</strong>m of the required solution. Once the algebraic equation has been solved<strong>for</strong> this Laplace trans<strong>for</strong>m, the general solution to the original ODE can beobtained by per<strong>for</strong>ming an inverse Laplace trans<strong>for</strong>m. One advantage of thismethod is that, <strong>for</strong> given boundary conditions, it provides the solution in justone step, instead of having to find the complementary function <strong>and</strong> particularintegral separately.In order to apply the method we need only two results from Laplace trans<strong>for</strong>mtheory (see section 13.2). First, the Laplace trans<strong>for</strong>m of a function f(x) is definedby¯f(s) ≡∫ ∞0e −sx f(x) dx, (15.31)from which we can derive the second useful relation. This concerns the Laplacetrans<strong>for</strong>m of the nth derivative of f(x):f (n) (s) =s n¯f(s) − s n−1 f(0) − s n−2 f ′ (0) − ···− sf (n−2) (0) − f (n−1) (0),(15.32)where the primes <strong>and</strong> superscripts in parentheses denote differentiation withrespect to x. Using these relations, along with table 13.1, on p. 455, which givesLaplace trans<strong>for</strong>ms of st<strong>and</strong>ard functions, we are in a position to solve a linearODE with constant coefficients by this method.501


HIGHER-ORDER ORDINARY DIFFERENTIAL EQUATIONS◮Solved 2 ydx − 3 dy2 dx +2y =2e−x , (15.33)subject to the boundary conditions y(0) = 2, y ′ (0) = 1.Taking the Laplace trans<strong>for</strong>m of (15.33) <strong>and</strong> using the table of st<strong>and</strong>ard results we obtains 2 ȳ(s) − sy(0) − y ′ (0) − 3 [sȳ(s) − y(0)] +2ȳ(s) = 2s +1 ,which reduces to(s 2 − 3s +2)ȳ(s) − 2s +5= 2s +1 . (15.34)Solving this algebraic equation <strong>for</strong> ȳ(s), the Laplace trans<strong>for</strong>m of the required solution to(15.33), we obtain2s 2 − 3s − 3ȳ(s) =(s +1)(s − 1)(s − 2) = 13(s +1) + 2s − 1 − 13(s − 2) , (15.35)where in the final step we have used partial fractions. Taking the inverse Laplace trans<strong>for</strong>mof (15.35), again using table 13.1, we find the specific solution to (15.33) to bey(x) = 1 3 e−x +2e x − 1 3 e2x . ◭Note that if the boundary conditions in a problem are given as symbols, ratherthan just numbers, then the step involving partial fractions can often involvea considerable amount of algebra. The Laplace trans<strong>for</strong>m method is also veryconvenient <strong>for</strong> solving sets of simultaneous linear ODEs with constant coefficients.◮Two electrical circuits, both of negligible resistance, each consist of a coil having selfinductanceL <strong>and</strong> a capacitor having capacitance C. The mutual inductance of the twocircuits is M. There is no source of e.m.f. in either circuit. Initially the second capacitoris given a charge CV 0 , the first capacitor being uncharged, <strong>and</strong> at time t =0aswitchinthe second circuit is closed to complete the circuit. Find the subsequent current in the firstcircuit.Subject to the initial conditions q 1 (0) = ˙q 1 (0) = ˙q 2 (0) = 0 <strong>and</strong> q 2 (0) = CV 0 = V 0 /G, say,we have to solveL¨q 1 + M¨q 2 + Gq 1 =0,M¨q 1 + L¨q 2 + Gq 2 =0.On taking the Laplace trans<strong>for</strong>m of the above equations, we obtain(Ls 2 + G)¯q 1 + Ms 2¯q 2 = sMV 0 C,Ms 2¯q 1 +(Ls 2 + G)¯q 2 = sLV 0 C.Eliminating ¯q 2 <strong>and</strong> rewriting as an equation <strong>for</strong> ¯q 1 , we findMV 0 s¯q 1 (s) =[(L + M)s 2 + G ][(L − M)s 2 + G ]= V []0 (L + M)s (L − M)s− .2G (L + M)s 2 + G (L − M)s 2 + G502


15.2 LINEAR EQUATIONS WITH VARIABLE COEFFICIENTSUsing table 13.1,q 1 (t) = 1 V 2 0C(cos ω 1 t − cos ω 2 t),where ω1 2(L + M) =G <strong>and</strong> ω2 2 (L − M) =G. Thus the current is given byi 1 (t) = 1 V 2 0C(ω 2 sin ω 2 t − ω 1 sin ω 1 t). ◭Solution method. Per<strong>for</strong>m a Laplace trans<strong>for</strong>m, as defined in (15.31), on the entireequation, using (15.32) to calculate the trans<strong>for</strong>m of the derivatives. Then solve theresulting algebraic equation <strong>for</strong> ȳ(s), the Laplace trans<strong>for</strong>m of the required solutionto the ODE. By using the method of partial fractions <strong>and</strong> consulting a table ofLaplace trans<strong>for</strong>ms of st<strong>and</strong>ard functions, calculate the inverse Laplace trans<strong>for</strong>m.The resulting function y(x) is the solution of the ODE that obeys the given boundaryconditions.15.2 Linear equations with variable coefficientsThere is no generally applicable method of solving equations with coefficientsthat are functions of x. Nevertheless, there are certain cases in which a solution ispossible. Some of the methods discussed in this section are also useful in findingthe general solution or particular integral <strong>for</strong> equations with constant coefficientsthat have proved impenetrable by the techniques discussed above.15.2.1 The Legendre <strong>and</strong> Euler linear equationsLegendre’s linear equation has the <strong>for</strong>ma n (αx + β) n dn ydx n + ···+ a 1(αx + β) dydx + a 0y = f(x), (15.36)where α, β <strong>and</strong> the a n are constants <strong>and</strong> may be solved by making the substitutionαx + β = e t . We then havedydx = dt dydx dt = α dyαx + β dtd 2 ydx 2 = ddxdydx = α 2 ( d 2 y(αx + β) 2 dt 2 − dy )dt<strong>and</strong> so on <strong>for</strong> higher derivatives. There<strong>for</strong>e we can write the terms of (15.36) as(αx + β) dydx = αdy dt ,(αx + β) 2 d2 ydx 2 = α2 d dt.(αx + β) n dn ydx n = αn d dt( ddt − 1 )y,( ) ( )d ddt − 1 ···dt − n +1 y.503(15.37)


HIGHER-ORDER ORDINARY DIFFERENTIAL EQUATIONSSubstituting equations (15.37) into the original equation (15.36), the latter becomesa linear ODE with constant coefficients, i.e.a n α n d ( ) ( )d ddt dt − 1 ···dt − n +1 y + ···+ a 1 α dy ( e t )dt + a − β0y = f ,αwhich can be solved by the methods of section 15.1.A special case of Legendre’s linear equation, <strong>for</strong> which α = 1 <strong>and</strong> β =0,isEuler’s equation,a n x n dn ydx n + ···+ a 1x dydx + a 0y = f(x); (15.38)it may be solved in a similar manner to the above by substituting x = e t .Iff(x) = 0 in (15.38) then substituting y = x λ leads to a simple algebraic equationin λ, which can be solved to yield the solution to (15.38). In the event that thealgebraic equation <strong>for</strong> λ has repeated roots, extra care is needed. If λ 1 is a k-foldroot (k >1) then the k linearly independent solutions corresponding to this rootare x λ1 ,x λ1 ln x,...,x λ1 (ln x) k−1 .◮Solvex 2 d2 ydx + x dy − 4y = 0 (15.39)2 dxby both of the methods discussed above.First we make the substitution x = e t , which, after cancelling e t , gives an equation withconstant coefficients, i.e.( )d ddt dt − 1 y + dydt − 4y =0 ⇒ d2 y− 4y =0. (15.40)dt2 Using the methods of section 15.1, the general solution of (15.40), <strong>and</strong> there<strong>for</strong>e of (15.39),is given byy = c 1 e 2t + c 2 e −2t = c 1 x 2 + c 2 x −2 .Since the RHS of (15.39) is zero, we can reach the same solution by substituting y = x λinto (15.39). This givesλ(λ − 1)x λ + λx λ − 4x λ =0,which reduces to(λ 2 − 4)x λ =0.This has the solutions λ = ±2, so we obtain again the general solutiony = c 1 x 2 + c 2 x −2 . ◭Solution method. If the ODE is of the Legendre <strong>for</strong>m (15.36) then substitute αx +β = e t . This results in an equation of the same order but with constant coefficients,which can be solved by the methods of section 15.1. If the ODE is of the Euler<strong>for</strong>m (15.38) with a non-zero RHS then substitute x = e t ; this again leads to anequation of the same order but with constant coefficients. If, however, f(x) =0inthe Euler equation (15.38) then the equation may also be solved by substituting504


15.2 LINEAR EQUATIONS WITH VARIABLE COEFFICIENTSy = x λ . This leads to an algebraic equation whose solution gives the allowed valuesof λ; the general solution is then the linear superposition of these functions.15.2.2 Exact equationsSometimes an ODE may be merely the derivative of another ODE of one orderlower. If this is the case then the ODE is called exact. The nth-order linear ODEa n (x) dn ydx n + ···+ a 1(x) dydx + a 0(x)y = f(x), (15.41)is exact if the LHS can be written as a simple derivative, i.e. ifa n (x) dn ydx n + ···+ a 0(x)y = d []b n−1 (x) dn−1 ydx dx n−1 + ···+ b 0(x)y . (15.42)It may be shown that, <strong>for</strong> (15.42) to hold, we requirea 0 (x) − a ′ 1(x)+a ′′2(x) − ···+(−1) n a n (n) (x) =0, (15.43)where the prime again denotes differentiation with respect to x. If (15.43) issatisfied then straight<strong>for</strong>ward integration leads to a new equation of one orderlower. If this simpler equation can be solved then a solution to the originalequation is obtained. Of course, if the above process leads to an equation that isitself exact then the analysis can be repeated to reduce the order still further.◮Solve(1 − x 2 ) d2 y dy− 3x − y =1. (15.44)dx2 dxComparing with (15.41), we have a 2 =1− x 2 , a 1 = −3x <strong>and</strong> a 0 = −1. It is easily shownthat a 0 − a ′ 1 + a′′ 2 = 0, so (15.44) is exact <strong>and</strong> can there<strong>for</strong>e be written in the <strong>for</strong>m[db 1 (x) dy ]dx dx + b 0(x)y =1. (15.45)Exp<strong>and</strong>ing the LHS of (15.45) we find( )d dyb 1dx dx + b d 2 y0y = b 1dx 2 +(b′ 1 + b 0 ) dydx + b′ 0y. (15.46)Comparing (15.44) <strong>and</strong> (15.46) we findb 1 =1− x 2 , b ′ 1 + b 0 = −3x, b ′ 0 = −1.These relations integrate consistently to give b 1 =1− x 2 <strong>and</strong> b 0 = −x, so (15.44) can bewritten as[d(1 − x 2 ) dy ]dx dx − xy =1. (15.47)Integrating (15.47) gives us directly the first-order linear ODEdy(dx − x)y = x + c 11 − x 2 1 − x , 2which can be solved by the method of subsection 14.2.4 <strong>and</strong> has the solutiony = c 1 sin −1 x + c 2√1 − x2− 1. ◭505


HIGHER-ORDER ORDINARY DIFFERENTIAL EQUATIONSIt is worth noting that, even if a higher-order ODE is not exact in its given <strong>for</strong>m,it may sometimes be made exact by multiplying through by some suitable function,an integrating factor, cf. subsection 14.2.3. Un<strong>for</strong>tunately, no straight<strong>for</strong>wardmethod <strong>for</strong> finding an integrating factor exists <strong>and</strong> one often has to rely oninspection or experience.◮Solvex(1 − x 2 ) d2 y dy− 3x2 − xy = x. (15.48)dx2 dxIt is easily shown that (15.48) is not exact, but we also see immediately that by multiplyingit through by 1/x we recover (15.44), which is exact <strong>and</strong> is solved above. ◭Another important point is that an ODE need not be linear to be exact,although no simple rule such as (15.43) exists if it is not linear. Nevertheless, it isoften worth exploring the possibility that a non-linear equation is exact, since itcould then be reduced in order by one <strong>and</strong> may lead to a soluble equation. Thisis discussed further in subsection 15.3.3.Solution method. For a linear ODE of the <strong>for</strong>m (15.41) check whether it is exactusing equation (15.43). If it is not then attempt to find an integrating factor whichwhen multiplying the equation makes it exact. Once the equation is exact write theLHS as a derivative as in (15.42) <strong>and</strong>, by exp<strong>and</strong>ing this derivative <strong>and</strong> comparingwith the LHS of the ODE, determine the functions b m (x) in (15.42). Integrate theresulting equation to yield another ODE, of one order lower. This may be solved orsimplified further if the new ODE is itself exact or can be made so.15.2.3 Partially known complementary functionSuppose we wish to solve the nth-order linear ODEa n (x) dn ydx n + ···+ a 1(x) dydx + a 0(x)y = f(x), (15.49)<strong>and</strong>wehappentoknowthatu(x) is a solution of (15.49) when the RHS isset to zero, i.e. u(x) is one part of the complementary function. By making thesubstitution y(x) =u(x)v(x), we can trans<strong>for</strong>m (15.49) into an equation of ordern − 1indv/dx. This simpler equation may prove soluble.In particular, if the original equation is of second order then we obtaina first-order equation in dv/dx, which may be soluble using the methods ofsection 14.2. In this way both the remaining term in the complementary function<strong>and</strong> the particular integral are found. This method there<strong>for</strong>e provides a usefulway of calculating particular integrals <strong>for</strong> second-order equations with variable(or constant) coefficients.506


15.2 LINEAR EQUATIONS WITH VARIABLE COEFFICIENTS◮Solved 2 y+ y =cosecx. (15.50)dx2 We see that the RHS does not fall into any of the categories listed in subsection 15.1.2,<strong>and</strong> so we are at an initial loss as to how to find the particular integral. However, thecomplementary function of (15.50) isy c (x) =c 1 sin x + c 2 cos x,<strong>and</strong> so let us choose the solution u(x) =cosx (we could equally well choose sin x) <strong>and</strong>make the substitution y(x) =v(x)u(x) =v(x)cosx into (15.50). This givescos x d2 v dv− 2sinx =cosecx, (15.51)dx2 dxwhich is a first-order linear ODE in dv/dx <strong>and</strong> may be solved by multiplying through bya suitable integrating factor, as discussed in subsection 14.2.4. Writing (15.51) asd 2 v dv− 2tanxdx2 dx = cosec xcos x , (15.52)we see that the required integrating factor is given by{ ∫ }exp −2 tan xdx =exp[2ln(cosx)] =cos 2 x.Multiplying both sides of (15.52) by the integrating factor cos 2 x we obtain(dcos 2 x dv )=cotx,dx dxwhich integrates to givecos 2 x dvdx =ln(sinx)+c 1.After rearranging <strong>and</strong> integrating again, this becomes∫∫v = sec 2 x ln(sin x) dx + c 1 sec 2 xdx=tanx ln(sin x) − x + c 1 tan x + c 2 .There<strong>for</strong>e the general solution to (15.50) is given by y = uv = v cos x, i.e.y = c 1 sin x + c 2 cos x +sinx ln(sin x) − x cos x,which contains the full complementary function <strong>and</strong> the particular integral. ◭Solution method. If u(x) is a known solution of the nth-order equation (15.49) withf(x) =0, then make the substitution y(x) =u(x)v(x) in (15.49). This leads to anequation of order n − 1 in dv/dx, which might be soluble.507


HIGHER-ORDER ORDINARY DIFFERENTIAL EQUATIONS15.2.4 Variation of parametersThe method of variation of parameters proves useful in finding particular integrals<strong>for</strong> linear ODEs with variable (<strong>and</strong> constant) coefficients. However, it requiresknowledge of the entire complementary function, not just of one part of it as inthe previous subsection.Suppose we wish to find a particular integral of the equationa n (x) dn ydx n + ···+ a 1(x) dydx + a 0(x)y = f(x), (15.53)<strong>and</strong> the complementary function y c (x) (the general solution of (15.53) withf(x) = 0) isy c (x) =c 1 y 1 (x)+c 2 y 2 (x)+···+ c n y n (x),where the functions y m (x) are known. We now assume that a particular integral of(15.53) can be expressed in a <strong>for</strong>m similar to that of the complementary function,but with the constants c m replaced by functions of x, i.e.weassumeaparticularintegral of the <strong>for</strong>my p (x) =k 1 (x)y 1 (x)+k 2 (x)y 2 (x)+···+ k n (x)y n (x). (15.54)This will no longer satisfy the complementary equation (i.e. (15.53) with the RHSset to zero) but might, with suitable choices of the functions k i (x), be made equalto f(x), thus producing not a complementary function but a particular integral.Since we have n arbitrary functions k 1 (x),k 2 (x),...,k n (x), but only one restrictionon them (namely the ODE), we may impose a further n − 1 constraints. Wecan choose these constraints to be as convenient as possible, <strong>and</strong> the simplestchoice is given byk 1(x)y ′ 1 (x)+k 2(x)y ′ 2 (x)+···+ k n(x)y ′ n (x) =0k 1(x)y ′ 1(x)+k ′ 2(x)y ′ 2(x)+···+ ′ k n(x)y ′ n(x) ′ =0.. (15.55)k 1(x)y ′ (n−2)1(x)+k 2(x)y ′ (n−2)2(x)+···+ k n(x)y ′ n (n−2) (x) =0k 1(x)y ′ (n−1)1(x)+k 2(x)y ′ (n−1)2(x)+···+ k n(x)y ′ n(n−1) (x) = f(x)a n (x) ,where the primes denote differentiation with respect to x. The last of theseequations is not a freely chosen constraint; given the previous n − 1 constraints<strong>and</strong> the original ODE, it must be satisfied.This choice of constraints is easily justified (although the algebra is quitemessy). Differentiating (15.54) with respect to x, we obtainy ′ p = k 1 y ′ 1 + k 2 y ′ 2 + ···+ k n y ′ n +[k ′ 1y 1 + k ′ 2y 2 + ···+ k ′ ny n ],where, <strong>for</strong> the moment, we drop the explicit x-dependence of these functions. Since508


15.2 LINEAR EQUATIONS WITH VARIABLE COEFFICIENTSwe are free to choose our constraints as we wish, let us define the expression inparentheses to be zero, giving the first equation in (15.55). Differentiating againwe findy p ′′ = k 1 y 1 ′′ + k 2 y 2 ′′ + ···+ k n y n ′′ +[k 1y ′ 1 ′ + k 2y ′ 2 ′ + ···+ k ny ′ n ′ ].Once more we can choose the expression in brackets to be zero, giving the secondequation in (15.55). We can repeat this procedure, choosing the correspondingexpression in each case to be zero. This yields the first n − 1 equations in (15.55).The mth derivative of y p <strong>for</strong> m


HIGHER-ORDER ORDINARY DIFFERENTIAL EQUATIONSintegration equal to zero, to give k m (x). The general solution to (15.53) is thengiven byn∑y(x) =y c (x)+y p (x) = [c m + k m (x)]y m (x).Note that if the constants of integration are included in the k m (x) then, as wellas finding the particular integral, we redefine the arbitrary constants c m in thecomplementary function.◮Use the variation-of-parameters method to solved 2 y+ y =cosecx, (15.57)dx2 subject to the boundary conditions y(0) = y(π/2) = 0.The complementary function of (15.57) is againy c (x) =c 1 sin x + c 2 cos x.We there<strong>for</strong>e assume a particular integral of the <strong>for</strong>my p (x) =k 1 (x)sinx + k 2 (x)cosx,<strong>and</strong> impose the additional constraints of (15.55), i.e.k 1(x)sinx ′ + k 2(x)cosx ′ =0,k 1(x)cosx ′ − k 2(x)sinx ′ =cosecx.Solving these equations <strong>for</strong> k 1 ′ (x) <strong>and</strong>k′ 2 (x) givesk 1(x) ′ =cosx cosec x =cotx,k 2(x) ′ =− sin x cosec x = −1.Hence, ignoring the constants of integration, k 1 (x) <strong>and</strong>k 2 (x) are given byk 1 (x) = ln(sin x),k 2 (x) =−x.The general solution to the ODE (15.57) is there<strong>for</strong>ey(x) =[c 1 +ln(sinx)] sin x +(c 2 − x)cosx,which is identical to the solution found in subsection 15.2.3. Applying the boundaryconditions y(0) = y(π/2) = 0 we find c 1 = c 2 =0<strong>and</strong>soy(x) = ln(sin x)sinx − x cos x. ◭m=1Solution method. If the complementary function of (15.53) is known then assumea particular integral of the same <strong>for</strong>m but with the constants replaced by functionsof x. Impose the constraints in (15.55) <strong>and</strong> solve the resulting system of equations<strong>for</strong> the unknowns k ′ 1 (x),k′ 2 ,...,k′ n(x). Integrate these functions, setting constants ofintegration equal to zero, to obtain k 1 (x),k 2 (x),...,k n (x) <strong>and</strong> hence the particularintegral.510


15.2 LINEAR EQUATIONS WITH VARIABLE COEFFICIENTS15.2.5 Green’s functionsThe Green’s function method of solving linear ODEs bears a striking resemblanceto the method of variation of parameters discussed in the previous subsection;it too requires knowledge of the entire complementary function in order to findthe particular integral <strong>and</strong> there<strong>for</strong>e the general solution. The Green’s functionapproach differs, however, since once the Green’s function <strong>for</strong> a particular LHSof (15.1) <strong>and</strong> particular boundary conditions has been found, then the solution<strong>for</strong> any RHS (i.e. any f(x)) can be written down immediately, albeit in the <strong>for</strong>mof an integral.Although the Green’s function method can be approached by considering thesuperposition of eigenfunctions of the equation (see chapter 17) <strong>and</strong> is alsoapplicable to the solution of partial differential equations (see chapter 21), thissection adopts a more utilitarian approach based on the properties of the Diracdelta function (see subsection 13.1.3) <strong>and</strong> deals only with the use of Green’sfunctions in solving ODEs.Let us again consider the equationa n (x) dn ydx n + ···+ a 1(x) dydx + a 0(x)y = f(x), (15.58)but <strong>for</strong> the sake of brevity we now denote the LHS by Ly(x), i.e. as a lineardifferential operator acting on y(x). Thus (15.58) now readsLy(x) =f(x). (15.59)Let us suppose that a function G(x, z) (the Green’s function) exists such that thegeneral solution to (15.59), which obeys some set of imposed boundary conditionsin the range a ≤ x ≤ b, is given byy(x) =∫ baG(x, z)f(z) dz, (15.60)where z is the integration variable. If we apply the linear differential operator Lto both sides of (15.60) <strong>and</strong> use (15.59) then we obtainLy(x) =∫ ba[LG(x, z)] f(z) dz = f(x). (15.61)Comparison of (15.61) with a st<strong>and</strong>ard property of the Dirac delta function (seesubsection 13.1.3), namelyf(x) =∫ baδ(x − z)f(z) dz,<strong>for</strong> a ≤ x ≤ b, shows that <strong>for</strong> (15.61) to hold <strong>for</strong> any arbitrary function f(x), werequire (<strong>for</strong> a ≤ x ≤ b) thatLG(x, z) =δ(x − z), (15.62)511


HIGHER-ORDER ORDINARY DIFFERENTIAL EQUATIONSi.e. the Green’s function G(x, z) must satisfy the original ODE with the RHS setequal to a delta function. G(x, z) may be thought of physically as the response ofa system to a unit impulse at x = z.In addition to (15.62), we must impose two further sets of restrictions onG(x, z). The first is the requirement that the general solution y(x) in (15.60) obeysthe boundary conditions. For homogeneous boundary conditions, in which y(x)<strong>and</strong>/or its derivatives are required to be zero at specified points, this is mostsimply arranged by dem<strong>and</strong>ing that G(x, z) itself obeys the boundary conditionswhen it is considered as a function of x alone; if, <strong>for</strong> example, we requirey(a) =y(b) = 0 then we should also dem<strong>and</strong> G(a, z) =G(b, z) = 0. Problemshaving inhomogeneous boundary conditions are discussed at the end of thissubsection.The second set of restrictions concerns the continuity or discontinuity of G(x, z)<strong>and</strong> its derivatives at x = z <strong>and</strong> can be found by integrating (15.62) with respectto x over the small interval [z − ɛ, z + ɛ] <strong>and</strong> taking the limit as ɛ → 0. We thenobtainlimn∑∫ z+ɛɛ→0m=0z−ɛa m (x) dm G(x, z)dx m dx = lim∫ z+ɛɛ→0 z−ɛδ(x − z) dx =1. (15.63)Since d n G/dx n exists at x = z but with value infinity, the (n−1)th-order derivativemust have a finite discontinuity there, whereas all the lower-order derivatives,d m G/dx m <strong>for</strong> m


15.2 LINEAR EQUATIONS WITH VARIABLE COEFFICIENTS(ii) When considered as a function of x alone G(x, z) obeys the specified(homogeneous) boundary conditions on y(x).(iii) The derivatives of G(x, z) with respect to x up to order n−2 are continuousat x = z, but the (n − 1)th-order derivative has a discontinuity of 1/a n (z)at this point.◮Use Green’s functions to solved 2 y+ y =cosecx, (15.66)dx2 subject to the boundary conditions y(0) = y(π/2) = 0.From (15.62) we see that the Green’s function G(x, z) mustsatisfyd 2 G(x, z)+ G(x, z) =δ(x − z). (15.67)dx 2Now it is clear that <strong>for</strong> x ≠ z the RHS of (15.67) is zero, <strong>and</strong> we are left with thetask of finding the general solution to the homogeneous equation, i.e. the complementaryfunction. The complementary function of (15.67) consists of a linear superposition of sin x<strong>and</strong> cos x <strong>and</strong> must consist of different superpositions on either side of x = z, since its(n − 1)th derivative (i.e. the first derivative in this case) is required to have a discontinuitythere. There<strong>for</strong>e we assume the <strong>for</strong>m of the Green’s function to be{ A(z)sinx + B(z)cosx <strong>for</strong> xz.Note that we have per<strong>for</strong>med a similar (but not identical) operation to that used in thevariation-of-parameters method, i.e. we have replaced the constants in the complementaryfunction with functions (this time of z).We must now impose the relevant restrictions on G(x, z) in order to determine thefunctions A(z),...,D(z). The first of these is that G(x, z) should itself obey the homogeneousboundary conditions G(0,z)=G(π/2,z) = 0. This leads to the conclusion that B(z) =C(z) = 0, so we now have{ A(z)sinx <strong>for</strong> xz.The second restriction is the continuity conditions given in equations (15.64), (15.65),namely that, <strong>for</strong> this second-order equation, G(x, z) is continuous at x = z <strong>and</strong> dG/dx hasa discontinuity of 1/a 2 (z) = 1 at this point. Applying these two constraints we haveD(z)cosz − A(z)sinz =0−D(z)sinz − A(z)cosz =1.Solving these equations <strong>for</strong> A(z) <strong>and</strong>D(z), we findA(z) =− cos z, D(z) =− sin z.Thus we have{ − cos z sin x <strong>for</strong> xz.There<strong>for</strong>e, from (15.60), the general solution to (15.66) that obeys the boundary conditions513


HIGHER-ORDER ORDINARY DIFFERENTIAL EQUATIONSy(0) = y(π/2) = 0 is given byy(x) =∫ π/20= − cos xG(x, z)coseczdz∫ x0sin z cosec zdz− sin x= −x cos x +sinx ln(sin x),∫ π/2xcos z cosec zdzwhich agrees with the result obtained in the previous subsections. ◭As mentioned earlier, once a Green’s function has been obtained <strong>for</strong> a givenLHS <strong>and</strong> boundary conditions, it can be used to find a general solution <strong>for</strong> anyRHS; thus, the solution of d 2 y/dx 2 + y = f(x), with y(0) = y(π/2) = 0, is givenimmediately byy(x) =∫ π/20= − cos xG(x, z)f(z) dz∫ x0sin zf(z) dz − sin x∫ π/2xcos zf(z) dz. (15.68)As an example, the reader may wish to verify that if f(x) =sin2x then (15.68)gives y(x) =(− sin 2x)/3, a solution easily verified by direct substitution. Ingeneral, analytic integration of (15.68) <strong>for</strong> arbitrary f(x) will prove intractable;then the integrals must be evaluated numerically.Another important point is that although the Green’s function method abovehas provided a general solution, it is also useful <strong>for</strong> finding a particular integralif the complementary function is known. This is easily seen since in (15.68) theconstant integration limits 0 <strong>and</strong> π/2 lead merely to constant values by whichthe factors sin x <strong>and</strong> cos x are multiplied; thus the complementary function isreconstructed. The rest of the general solution, i.e. the particular integral, comesto − ∫ x ,<strong>and</strong>sodropping the constant integration limits, we can find just the particular integral.For example, a particular integral of d 2 y/dx 2 + y = f(x) that satisfies the aboveboundary conditions is given byfrom the variable integration limit x. There<strong>for</strong>e by changing ∫ π/2xy p (x) =− cos x∫ xsin zf(z) dz +sinx∫ xcos zf(z) dz.A very important point to realise about the Green’s function method is that aparticular G(x, z) applies to a given LHS of an ODE <strong>and</strong> the imposed boundaryconditions, i.e. the same equation with different boundary conditions will have adifferent Green’s function. To illustrate this point, let us consider again the ODEsolved in (15.68), but with different boundary conditions.514


15.2 LINEAR EQUATIONS WITH VARIABLE COEFFICIENTS◮Use Green’s functions to solved 2 y+ y = f(x), (15.69)dx2 subject to the one-point boundary conditions y(0) = y ′ (0) = 0.We again require (15.67) to hold <strong>and</strong> so again we assume a Green’s function of the <strong>for</strong>m{ A(z)sinx + B(z)cosx <strong>for</strong> xz.However, we now require G(x, z) to obey the boundary conditions G(0,z)=G ′ (0,z)=0,which imply A(z) =B(z) = 0. There<strong>for</strong>e we have{ 0 <strong>for</strong> xz.Applying the continuity conditions on G(x, z) as be<strong>for</strong>e now givesC(z)sinz + D(z)cosz =0,C(z)cosz − D(z)sinz =1,which are solved to giveC(z) =cosz, D(z) =− sin z.So finally the Green’s function is given by{ 0 <strong>for</strong> xz,<strong>and</strong> the general solution to (15.69) that obeys the boundary conditions y(0) = y ′ (0) = 0 isy(x) ==∫ ∞0∫ x0G(x, z)f(z) dzsin(x − z)f(z) dz. ◭Finally, we consider how to deal with inhomogeneous boundary conditionssuch as y(a) =α, y(b) =β or y(0) = y ′ (0) = γ, whereα, β, γ are non-zero. Thesimplest method of solution in this case is to make a change of variable such thatthe boundary conditions in the new variable, u say, are homogeneous, i.e. u(a) =u(b) =0oru(0) = u ′ (0) = 0 etc. For nth-order equations we generally requiren boundary conditions to fix the solution, but these n boundary conditions canbe of various types: we could have the n-point boundary conditions y(x m )=y m<strong>for</strong> m =1ton, or the one-point boundary conditions y(x 0 )=y ′ (x 0 )=··· =y (n−1) (x 0 )=y 0 , or something in between. In all cases a suitable change of variableisu = y − h(x),where h(x) isan(n − 1)th-order polynomial that obeys the boundary conditions.515


HIGHER-ORDER ORDINARY DIFFERENTIAL EQUATIONSFor example, if we consider the second-order case with boundary conditionsy(a) =α, y(b) =β then a suitable change of variable isu = y − (mx + c),where y = mx + c is the straight line through the points (a, α) <strong>and</strong>(b, β), <strong>for</strong> whichm =(α − β)/(a − b) <strong>and</strong>c =(βa − αb)/(a − b). Alternatively, if the boundaryconditions <strong>for</strong> our second-order equation are y(0) = y ′ (0) = γ then we wouldmake the same change of variable, but this time y = mx + c would be the straightline through (0,γ) with slope γ, i.e.m = c = γ.Solution method. Require that the Green’s function G(x, z) obeys the original ODE,but with the RHS set to a delta function δ(x − z). This is equivalent to assumingthat G(x, z) is given by the complementary function of the original ODE, with theconstants replaced by functions of z; these functions are different <strong>for</strong> xz. Now require also that G(x, z) obeys the given homogeneous boundary conditions<strong>and</strong> impose the continuity conditions given in (15.64) <strong>and</strong> (15.65). The generalsolution to the original ODE is then given by (15.60). For inhomogeneous boundaryconditions, make the change of dependent variable u = y − h(x), whereh(x) is apolynomial obeying the given boundary conditions.15.2.6 Canonical <strong>for</strong>m <strong>for</strong> second-order equationsIn this section we specialise from nth-order linear ODEs with variable coefficientsto those of order 2. In particular we consider the equationd 2 ydx 2 + a 1(x) dydx + a 0(x)y = f(x), (15.70)which has been rearranged so that the coefficient of d 2 y/dx 2 is unity. By makingthe substitution y(x) =u(x)v(x) we obtain( ) ( 2uv ′′ ′u+u + a 1 v ′ ′′ + a 1 u ′ )+ a 0 u+v = f uu , (15.71)where the prime denotes differentiation with respect to x. Since (15.71) would bemuch simplified if there were no term in v ′ ,letuschooseu(x) such that the firstfactor in parentheses on the LHS of (15.71) is zero, i.e.2u ′{ ∫u + a 1 =0 ⇒ u(x) =exp − 1 2We then obtain an equation of the <strong>for</strong>m}a 1 (z) dz . (15.72)d 2 v+ g(x)v = h(x), (15.73)dx2 516


15.2 LINEAR EQUATIONS WITH VARIABLE COEFFICIENTSwhereg(x) =a 0 (x) − 1 4 [a 1(x)] 2 − 1 2 a′ 1{ ∫ }(x)h(x) =f(x)exp a 1 (z) dz .Since (15.73) is of a simpler <strong>for</strong>m than the original equation, (15.70), it mayprove easier to solve.12◮Solve4x 2 d2 y dy+4xdx2 dx +(x2 − 1)y =0. (15.74)Dividing (15.74) through by 4x 2 , we see that it is of the <strong>for</strong>m (15.70) with a 1 (x) =1/x,a 0 (x) =(x 2 − 1)/4x 2 <strong>and</strong> f(x) = 0. There<strong>for</strong>e, making the substitution( ∫ ) 1y = vu = v exp −2x dx = √ Av , xwe obtaind 2 vdx + v =0. (15.75)2 4Equation (15.75) is easily solved to givev = c 1 sin 1 x + c 2 2 cos 1 x, 2so the solution of (15.74) isy = √ v = c 1 sin 1 x + c 2 2 cos 1 2√ x . ◭x xAs an alternative to choosing u(x) such that the coefficient of v ′ in (15.71) iszero, we could choose a different u(x) such that the coefficient of v vanishes. Forthis to be the case, we see from (15.71) that we would requireu ′′ + a 1 u ′ + a 0 u =0,so u(x) would have to be a solution of the original ODE with the RHS set tozero, i.e. part of the complementary function. If such a solution were known thenthe substitution y = uv would yield an equation with no term in v, which couldbe solved by two straight<strong>for</strong>ward integrations. This is a special (second-order)case of the method discussed in subsection 15.2.3.Solution method. Write the equation in the <strong>for</strong>m (15.70), then substitute y = uv,where u(x) is given by (15.72). This leads to an equation of the <strong>for</strong>m (15.73), inwhich there is no term in dv/dx <strong>and</strong> which may be easier to solve. Alternatively,if part of the complementary function is known then follow the method of subsection15.2.3.517


HIGHER-ORDER ORDINARY DIFFERENTIAL EQUATIONS15.3 General ordinary differential equationsIn this section, we discuss miscellaneous methods <strong>for</strong> simplifying general ODEs.These methods are applicable to both linear <strong>and</strong> non-linear equations <strong>and</strong> insome cases may lead to a solution. More often than not, however, finding aclosed-<strong>for</strong>m solution to a general non-linear ODE proves impossible.15.3.1 Dependent variable absentIf an ODE does not contain the dependent variable y explicitly, but only itsderivatives, then the change of variable p = dy/dx leads to an equation of oneorder lower.◮Solved 2 ydx +2dy =4x (15.76)2 dxThis is trans<strong>for</strong>med by the substitution p = dy/dx to the first-order equationdp+2p =4x. (15.77)dxThe solution to (15.77) is then found by the method of subsection 14.2.4 <strong>and</strong> readsp = dydx = ae−2x +2x − 1,where a is a constant. Thus by direct integration the solution to the original equation,(15.76), isy(x) =c 1 e −2x + x 2 − x + c 2 . ◭An extension to the above method is appropriate if an ODE contains onlyderivatives of y that are of order m <strong>and</strong> greater. Then the substitution p = d m y/dx mreduces the order of the ODE by m.Solution method. If the ODE contains only derivatives of y that are of order m <strong>and</strong>greater then the substitution p = d m y/dx m reduces the order of the equation by m.15.3.2 Independent variable absentIf an ODE does not contain the independent variable x explicitly, except in d/dx,d 2 /dx 2 etc., then as in the previous subsection we make the substitution p = dy/dx518


15.3 GENERAL ORDINARY DIFFERENTIAL EQUATIONSbut also writed 2 ydx 2 = dpdx = dy dpdx dy = p dpdyd 3 ydx 3 = d (p dp )= dy ddx dy dx dy(p dp )( ) 2= p 2 d2 p dpdy dy 2 + p , (15.78)dy<strong>and</strong> so on <strong>for</strong> higher-order derivatives. This leads to an equation of one orderlower.◮Solve( ) 21+y d2 y dydx + =0. (15.79)2 dxMaking the substitutions dy/dx = p <strong>and</strong> d 2 y/dx 2 = p(dp/dy) we obtain the first-orderODE1+yp dpdy + p2 =0,which is separable <strong>and</strong> may be solved as in subsection 14.2.1 to obtain(1 + p 2 )y 2 = c 1 .Using p = dy/dx we there<strong>for</strong>e have√p = dydx = ± c 2 1 − y2,y 2which may be integrated to give the general solution of (15.79); after squaring this reads(x + c 2 ) 2 + y 2 = c 2 1. ◭Solution method. If the ODE does not contain x explicitly then substitute p =dy/dx, along with the relations <strong>for</strong> higher derivatives given in (15.78), to obtain anequation of one order lower, which may prove easier to solve.15.3.3 Non-linear exact equationsAs discussed in subsection 15.2.2, an exact ODE is one that can be obtained bystraight<strong>for</strong>ward differentiation of an equation of one order lower. Moreover, thenotion of exact equations is useful <strong>for</strong> both linear <strong>and</strong> non-linear equations, sincean exact equation can be immediately integrated. It is possible, of course, thatthe resulting equation may itself be exact, so that the process can be repeated.In the non-linear case, however, there is no simple relation (such as (15.43) <strong>for</strong>the linear case) by which an equation can be shown to be exact. Nevertheless, ageneral procedure does exist <strong>and</strong> is illustrated in the following example.519


HIGHER-ORDER ORDINARY DIFFERENTIAL EQUATIONS◮Solve2y d3 ydx +6dy d 2 y= x. (15.80)3 dx dx2 Directing our attention to the term on the LHS of (15.80) that contains the highest-orderderivative, i.e. 2yd 3 y/dx 3 , we see that it can be obtained by differentiating 2yd 2 y/dx 2 since( )d2y d2 y=2y d3 ydx dx 2 dx +2dy d 2 y3 dx dx . (15.81)2Rewriting the LHS of (15.80) using (15.81), we are left with 4(dy/dx)(d 2 y/dy 2 ), which mayitself be written as a derivative, i.e.[4 dy d 2 ydx dx = d 22 dx( ) ] 2 dy. (15.82)dxSince, there<strong>for</strong>e, we can write the LHS of (15.80) as a sum of simple derivatives of otherfunctions, (15.80) is exact. Integrating (15.80) with respect to x, <strong>and</strong> using (15.81) <strong>and</strong>(15.82), now gives( ) 2 ∫2y d2 y dydx +2 = 2 dxxdx= x22 + c 1. (15.83)Now we can repeat the process to find whether (15.83) is itself exact. Considering the termon the LHS of (15.83) that contains the highest-order derivative, i.e. 2yd 2 y/dx 2 , we notethat we obtain this by differentiating 2ydy/dx, as follows:ddx(2y dydx)=2y d2 ydx 2 +2 ( dydxThe above expression already contains all the terms on the LHS of (15.83), so we canintegrate (15.83) to give2y dydx = x36 + c 1x + c 2 .Integrating once more we obtain the solutiony 2 = x424 + c 1x 22 + c 2x + c 3 . ◭It is worth noting that both linear equations (as discussed in subsection 15.2.2)<strong>and</strong> non-linear equations may sometimes be made exact by multiplying throughby an appropriate integrating factor. Although no general method exists <strong>for</strong>finding such a factor, one may sometimes be found by inspection or inspiredguesswork.Solution method. Rearrange the equation so that all the terms containing y or itsderivatives are on the LHS, then check to see whether the equation is exact byattempting to write the LHS as a simple derivative. If this is possible then theequation is exact <strong>and</strong> may be integrated directly to give an equation of one orderlower. If the new equation is itself exact the process can be repeated.) 2.520


15.3 GENERAL ORDINARY DIFFERENTIAL EQUATIONS15.3.4 Isobaric or homogeneous equationsIt is straight<strong>for</strong>ward to generalise the discussion of first-order isobaric equationsgiven in subsection 14.2.6 to equations of general order n. Annth-order isobaricequation is one in which every term can be made dimensionally consistent upongiving y <strong>and</strong> dy each a weight m, <strong>and</strong>x <strong>and</strong> dx each a weight 1. Then the nthderivative of y with respect to x, <strong>for</strong> example, would have dimensions m in y<strong>and</strong> −n in x. In the special case m = 1, <strong>for</strong> which the equation is dimensionallyconsistent, the equation is called homogeneous (not to be confused with linearequations with a zero RHS). If an equation is isobaric or homogeneous then thechange in dependent variable y = vx m (y = vx in the homogeneous case) followedby the change in independent variable x = e t leadstoanequationinwhichthenew independent variable t is absent except in the <strong>for</strong>m d/dt.◮Solvex 3 d2 ydx 2 − (x2 + xy) dydx +(y2 + xy) =0. (15.84)Assigning y <strong>and</strong> dy the weight m, <strong>and</strong>x <strong>and</strong> dx the weight 1, the weights of the five termson the LHS of (15.84) are, from left to right: m +1, m +1, 2m, 2m, m +1. For theseweights all to be equal we require m = 1; thus (15.84) is a homogeneous equation. Since itis homogeneous we now make the substitution y = vx, which, after dividing the resultingequation through by x 3 ,givesx d2 v dv+(1− v) =0. (15.85)dx2 dxNow substituting x = e t into (15.85) we obtain (after some working)d 2 vdt − v dv =0, (15.86)2 dtwhich can be integrated directly to givedvdt = 1 2 v2 + c 1 . (15.87)Equation (15.87) is separable, <strong>and</strong> integrates to give∫1t + d dv2 2 =v 2 + d 2 1= 1 ( ) vtan −1 .d 1 d 1Rearranging <strong>and</strong> using x = e t <strong>and</strong> y = vx we finally obtain the solution to (15.84) asy = d 1 x tan ( 1d )2 1 ln x + d 1 d 2 . ◭Solution method. Assume that y <strong>and</strong> dy have weight m, <strong>and</strong> x <strong>and</strong> dx weight 1,<strong>and</strong> write down the combined weights of each term in the ODE. If these weights canbe made equal by assuming a particular value <strong>for</strong> m then the equation is isobaric(or homogeneous if m =1). Making the substitution y = vx m followed by x = e tleads to an equation in which the new independent variable t is absent except in the<strong>for</strong>m d/dt.521


HIGHER-ORDER ORDINARY DIFFERENTIAL EQUATIONS15.3.5 Equations homogeneous in x or y aloneIt will be seen that the intermediate equation (15.85) in the example of theprevious subsection was simplified by the substitution x = e t , in that this led toan equation in which the new independent variable t occurred only in the <strong>for</strong>md/dt; see (15.86). A closer examination of (15.85) reveals that it is dimensionallyconsistent in the independent variable x taken alone; this is equivalent to givingthe dependent variable <strong>and</strong> its differential a weight m = 0. For any equation thatis homogeneous in x alone, the substitution x = e t will lead to an equation thatdoes not contain the new independent variable t except as d/dt. Note that theEuler equation of subsection 15.2.1 is a special, linear example of an equationhomogeneous in x alone. Similarly, if an equation is homogeneous in y alone, thensubstituting y = e v leads to an equation in which the new dependent variable, v,occurs only in the <strong>for</strong>m d/dv.◮Solvex 2 d2 ydx 2 + x dydx + 2 y 3 =0.This equation is homogeneous in x alone, <strong>and</strong> on substituting x = e t we obtaind 2 ydt + 2 2 y =0, 3which does not contain the new independent variable t except as d/dt. Such equationsmay often be solved by the method of subsection 15.3.2, but in this case we can integratedirectly to obtaindydt = √ 2(c 1 +1/y 2 ).This equation is separable, <strong>and</strong> we find∫dy√2(c1 +1/y 2 ) = t + c 2.By multiplying the numerator <strong>and</strong> denominator of the integr<strong>and</strong> on the LHS by y, we findthe solution√c1 y 2 +1√2c1= t + c 2 .Remembering that t =lnx, we finally obtain√c1 y 2 +1√ =lnx + c 2 . ◭2c1Solution method. If the weight of x taken alone is the same in every term in theODE then the substitution x = e t leads to an equation in which the new independentvariable t is absent except in the <strong>for</strong>m d/dt. If the weight of y taken alone is thesame in every term then the substitution y = e v leads to an equation in which thenew dependent variable v is absent except in the <strong>for</strong>m d/dv.522


15.4 EXERCISES15.3.6 Equations having y = Ae x as a solutionFinally, we note that if any general (linear or non-linear) nth-order ODE issatisfied identically by assuming thaty = dydx = ···= dn ydx n (15.88)then y = Ae x is a solution of that equation. This must be so because y = Ae x isa non-zero function that satisfies (15.88).◮Find a solution of(x 2 + x) dy d 2 ydx dx − 2 x2 y dy ( ) 2 dydx − x =0. (15.89)dxSetting y = dy/dx = d 2 y/dx 2 in (15.89), we obtain(x 2 + x)y 2 − x 2 y 2 − xy 2 =0,which is satisfied identically. There<strong>for</strong>e y = Ae x is a solution of (15.89); this is easilyverified by directly substituting y = Ae x into (15.89). ◭Solution method. If the equation is satisfied identically by making the substitutionsy = dy/dx = ···= d n y/dx n then y = Ae x is a solution.15.4 Exercises15.1 A simple harmonic oscillator, of mass m <strong>and</strong> natural frequency ω 0 , experiencesan oscillating driving <strong>for</strong>ce f(t) =ma cos ωt. There<strong>for</strong>e, its equation of motion isd 2 xdt + 2 ω2 0x = a cos ωt,where x is its position. Given that at t = 0 we have x = dx/dt = 0, find thefunction x(t). Describe the solution if ω is approximately, but not exactly, equalto ω 0 .15.2 Find the roots of the auxiliary equation <strong>for</strong> the following. Hence solve them <strong>for</strong>the boundary conditions stated.(a) d2 fdt +2df2 dt +5f =0, with f(0) = 1,f′ (0) = 0.(b) d2 fdt +2df2 dt +5f = e−t cos 3t, with f(0) = 0,f ′ (0) = 0.15.3 The theory of bent beams shows that at any point in the beam the ‘bendingmoment’ is given by K/ρ,whereK is a constant (that depends upon the beammaterial <strong>and</strong> cross-sectional shape) <strong>and</strong> ρ is the radius of curvature at that point.Consider a light beam of length L whose ends, x =0<strong>and</strong>x = L, are supportedat the same vertical height <strong>and</strong> which has a weight W suspended from its centre.Verify that at any point x (0 ≤ x ≤ L/2 <strong>for</strong> definiteness) the net magnitude ofthe bending moment (bending moment = <strong>for</strong>ce × perpendicular distance) due tothe weight <strong>and</strong> support reactions, evaluated on either side of x, isWx/2.523


HIGHER-ORDER ORDINARY DIFFERENTIAL EQUATIONSIf the beam is only slightly bent, so that (dy/dx) 2 ≪ 1, where y = y(x) isthedownward displacement of the beam at x, show that the beam profile satisfiesthe approximate equationd 2 ydx = − Wx2 2K .By integrating this equation twice <strong>and</strong> using physically imposed conditions onyour solution at x =0<strong>and</strong>x = L/2, show that the downward displacement atthe centre of the beam is WL 3 /(48K).15.4 Solve the differential equationd 2 fdt +6df2 dt +9f = e−t ,subject to the conditions f =0<strong>and</strong>df/dt = λ at t =0.Find the equation satisfied by the positions of the turning points of f(t) <strong>and</strong>hence, by drawing suitable sketch graphs, determine the number of turning pointsthe solution has in the range t>0if(a)λ =1/4, <strong>and</strong> (b) λ = −1/4.15.5 The function f(t) satisfies the differential equationd 2 fdt +8df2 dt +12f =12e−4t .For the following sets of boundary conditions determine whether it has solutions,<strong>and</strong>, if so, find them:(a) f(0) = 0, f ′ (0) = 0, f(ln √ 2) = 0;(b) f(0) = 0, f ′ (0) = −2, f(ln √ 2) = 0.15.6 Determine the values of α <strong>and</strong> β <strong>for</strong> which the following four functions arelinearly dependent:y 1 (x) =x cosh x +sinhx,y 2 (x) =x sinh x +coshx,y 3 (x) =(x + α)e x ,y 4 (x) =(x + β)e −x .You will find it convenient to work with those linear combinations of the y i (x)that can be written the most compactly.15.7 A solution of the differential equationd 2 ydx +2dy2 dx + y =4e−xtakes the value 1 when x = 0 <strong>and</strong> the value e −1 when x = 1. What is its valuewhen x =2?15.8 The two functions x(t) <strong>and</strong>y(t) satisfy the simultaneous equationsdx− 2y = − sin t,dtdy+2x =5cost.dtFind explicit expressions <strong>for</strong> x(t) <strong>and</strong>y(t), given that x(0) = 3 <strong>and</strong> y(0) = 2.Sketch the solution trajectory in the xy-plane <strong>for</strong> 0 ≤ t


15.4 EXERCISES15.9 Find the general solutions of(a)(b)d 3 y dy− 12 +16y =32x − 8,dx3 ( dx )(d 1 dy1+(2a coth 2ax)dx y dxydydx)=2a 2 ,where a is a constant.15.10 Use the method of Laplace trans<strong>for</strong>ms to solve(a)d 2 fdt +5df2 dt +6f =0, f(0) = 1, f′ (0) = −4,(b)d 2 fdt +2df2 dt +5f =0, f(0) = 1, f′ (0) = 0.15.11 The quantities x(t), y(t) satisfy the simultaneous equationsẍ +2nẋ + n 2 x =0,ÿ +2nẏ + n 2 y = µẋ,where x(0) = y(0) = ẏ(0) = 0 <strong>and</strong> ẋ(0) = λ. Show thaty(t) = ( 1 2 µλt2 1 − 1 nt) exp(−nt).315.12 Use Laplace trans<strong>for</strong>ms to solve, <strong>for</strong> t ≥ 0, the differential equationsẍ +2x + y =cost,ÿ +2x +3y =2cost,which describe a coupled system that starts from rest at the equilibrium position.Show that the subsequent motion takes place along a straight line in the xy-plane.Verify that the frequency at which the system is driven is equal to one of theresonance frequencies of the system; explain why there is no resonant behaviourin the solution you have obtained.15.13 Two unstable isotopes A <strong>and</strong> B <strong>and</strong> a stable isotope C have the following decayrates per atom present: A → B, 3s −1 ; A → C, 1s −1 ; B → C, 2s −1 . Initially aquantity x 0 of A is present, but there are no atoms of the other two types. UsingLaplace trans<strong>for</strong>ms, find the amount of C present at a later time t.15.14 For a lightly damped (γ < ω 0 ) harmonic oscillator driven at its undampedresonance frequency ω 0 , the displacement x(t) attimet satisfies the equationd 2 x dx+2γdt 2 dt + ω2 0x = F sin ω 0 t.Use Laplace trans<strong>for</strong>ms to find the displacement at a general time if the oscillatorstarts from rest at its equilibrium position.(a) Show that ultimately the oscillation has amplitude F/(2ω 0 γ), with a phaselag of π/2 relative to the driving <strong>for</strong>ce per unit mass F.(b) By differentiating the original equation, conclude that if x(t) is exp<strong>and</strong>ed asa power series in t <strong>for</strong> small t, then the first non-vanishing term is Fω 0 t 3 /6.Confirm this conclusion by exp<strong>and</strong>ing your explicit solution.15.15 The ‘golden mean’, which is said to describe the most aesthetically pleasingproportions <strong>for</strong> the sides of a rectangle (e.g. the ideal picture frame), is givenby the limiting value of the ratio of successive terms of the Fibonacci series u n ,which is generated byu n+2 = u n+1 + u n ,with u 0 =0<strong>and</strong>u 1 = 1. Find an expression <strong>for</strong> the general term of the series <strong>and</strong>525


HIGHER-ORDER ORDINARY DIFFERENTIAL EQUATIONSverify that the golden mean is equal to the larger root of the recurrence relation’scharacteristic equation.15.16 In a particular scheme <strong>for</strong> numerically modelling one-dimensional fluid flow, thesuccessive values, u n , of the solution are connected <strong>for</strong> n ≥ 1 by the differenceequationc(u n+1 − u n−1 )=d(u n+1 − 2u n + u n−1 ),where c <strong>and</strong> d are positive constants. The boundary conditions are u 0 =0<strong>and</strong>u M = 1. Find the solution to the equation, <strong>and</strong> show that successive values of u nwill have alternating signs if c>d.15.17 The first few terms of a series u n , starting with u 0 ,are1, 2, 2, 1, 6, −3. The seriesis generated by a recurrence relation of the <strong>for</strong>mu n = Pu n−2 + Qu n−4 ,where P <strong>and</strong> Q are constants. Find an expression <strong>for</strong> the general term of theseries <strong>and</strong> show that, in fact, the series consists of two interleaved series given byu 2m = 2 + 1 3 3 4m ,u 2m+1 = 7 − 1 3 3 4m ,<strong>for</strong> m =0, 1, 2,... .15.18 Find an explicit expression <strong>for</strong> the u n satisfyingu n+1 +5u n +6u n−1 =2 n ,given that u 0 = u 1 = 1. Deduce that 2 n − 26(−3) n is divisible by 5 <strong>for</strong> allnon-negative integers n.15.19 Find the general expression <strong>for</strong> the u n satisfyingu n+1 =2u n−2 − u nwith u 0 = u 1 =0<strong>and</strong>u 2 = 1, <strong>and</strong> show that they can be written in the <strong>for</strong>mu n = 1 ( )5 − 2n/2 3πn√ cos5 4 − φ ,where tan φ =2.15.20 Consider the seventh-order recurrence relationu n+7 − u n+6 − u n+5 + u n+4 − u n+3 + u n+2 + u n+1 − u n =0.Find the most general <strong>for</strong>m of its solution, <strong>and</strong> show that:(a) if only the four initial values u 0 =0,u 1 =2,u 2 =6<strong>and</strong>u 3 = 12, are specified,then the relation has one solution that cycles repeatedly through this set offour numbers;(b) but if, in addition, it is required that u 4 = 20, u 5 =30<strong>and</strong>u 6 = 42 then thesolution is unique, with u n = n(n +1).15.21 Find the general solution ofx 2 d2 ydx − x dy2 dx + y = x,given that y(1) = 1 <strong>and</strong> y(e) =2e.15.22 Find the general solution of(x +1) 2 d2 y+3(x +1)dydx2 dx + y = x2 .526


15.4 EXERCISES15.23 Prove that the general solution of(x − 2) d2 ydx 2 +3dy dx + 4yx 2 =0is given by[ (1 2y(x) = k(x − 2) 2 3x − 1 ) ]+ cx 2 .215.24 Use the method of variation of parameters to find the general solutions of(a)d 2 ydx 2 − y = xn ,(b) d2 ydx 2 − 2 dydx + y =2xex .15.25 Use the intermediate result of exercise 15.24(a) to find the Green’s function thatsatisfiesd 2 G(x, ξ)− G(x, ξ) =δ(x − ξ) with G(0,ξ)=G(1,ξ)=0.dx 215.26 Consider the equationF(x, y) =x(x +1) d2 ydx +(2− 2 x2 ) dy − (2 + x)y =0.dx(a) Given that y 1 (x) =1/x is one of its solutions, find a second linearly independentone,(i) by setting y 2 (x) =y 1 (x)u(x), <strong>and</strong>(ii) by noting the sum of the coefficients in the equation.(b) Hence, using the variation of parameters method, find the general solutionofF(x, y) =(x +1) 2 .15.27 Show generally that if y 1 (x) <strong>and</strong>y 2 (x) are linearly independent solutions ofd 2 y dy+ p(x) + q(x)y =0,dx2 dxwith y 1 (0) = 0 <strong>and</strong> y 2 (1) = 0, then the Green’s function G(x, ξ) <strong>for</strong> the interval0 ≤ x, ξ ≤ 1<strong>and</strong>withG(0,ξ)=G(1,ξ)=0canbewritteninthe<strong>for</strong>m{y1 (x)y 2 (ξ)/W (ξ) 0


HIGHER-ORDER ORDINARY DIFFERENTIAL EQUATIONS15.29 The equation of motion <strong>for</strong> a driven damped harmonic oscillator can be writtenẍ +2ẋ +(1+κ 2 )x = f(t),with κ ≠0.Ifitstartsfromrestwithx(0) = 0 <strong>and</strong> ẋ(0) = 0, find the correspondingGreen’s function G(t, τ) <strong>and</strong> verify that it can be written as a function of t − τonly. Find the explicit solution when the driving <strong>for</strong>ce is the unit step function,i.e. f(t) =H(t). Confirm your solution by taking the Laplace trans<strong>for</strong>ms of bothit <strong>and</strong> the original equation.15.30 Show that the Green’s function <strong>for</strong> the equationd 2 ydx + y 2 4 = f(x),subject to the boundary conditions y(0) = y(π) = 0, is given by{−2cos12G(x, z) =x sin 1 z 0 ≤ z ≤ x,2−2sin 1 x cos 1 z x ≤ z ≤ π.2 215.31 Find the Green’s function x = G(t, t 0 ) that solvesd 2 xdt + α dx2 dt = δ(t − t 0)under the initial conditions x = dx/dt =0att =0.Hencesolved 2 xdt + α dx2 dt = f(t),where f(t) =0<strong>for</strong>t0).15.32 Consider the equationd 2 y+ f(y) =0,dx2 where f(y) can be any function.(a) By multiplying through by dy/dx, obtain the general solution relating x <strong>and</strong>y.(b) A mass m, initially at rest at the point x = 0, is accelerated by a <strong>for</strong>ce[f(x) =A(x 0 − x) 1+2ln(1 − x )].x 0Its equation of motion is md 2 x/dt 2 = f(x). Find x as a function of time, <strong>and</strong>show that ultimately the particle has travelled a distance x 0 .15.33 Solve(2y d3 ydx +2 y +3 dy ) ( ) d 2 2y dy3 dx dx +2 = sinx.2 dx15.34 Find the general solution of the equationx d3 y ydx 3 +2d2 dx = Ax. 215.35 Express the equationd 2 y dy+4xdx2 dx +(4x2 +6)y = e −x2 sin 2xin canonical <strong>for</strong>m <strong>and</strong> hence find its general solution.528


15.5 HINTS AND ANSWERS15.36 Find the <strong>for</strong>m of the solutions of the equation( )dy d 3 y d 2 2 ( ) 2dx dx − 2 y dy+ =03 dx 2 dxthat have y(0) = ∞.[ You will need the result ∫ z cosech udu= − ln(cosech z +cothz). ]15.37 Consider the equation( ) 2x p y ′′ n +3− 2p p − 2+ x p−1 y ′ + x p−2 y = y n ,n − 1n − 1in which p ≠2<strong>and</strong>n>−1 but n ≠ 1. For the boundary conditions y(1) = 0 <strong>and</strong>y ′ (1) = λ, show that the solution is y(x) =v(x)x (p−2)/(n−1) ,wherev(x) isgivenby∫ v(x)0dz[λ2 +2z n+1 /(n +1) ] 1/2 =lnx.15.5 Hints <strong>and</strong> answers15.1 The function is a(ω0 2 − ω2 ) −1 (cos ωt − cos ω 0 t); <strong>for</strong> moderate t, x(t) is a sine waveof linearly increasing amplitude (t sin ω 0 t)/(2ω 0 ); <strong>for</strong> large t it shows beats ofmaximum amplitude 2(ω0 2 − ω2 ) −1 .15.3 Ignore the term y ′2 , compared with 1, in the expression <strong>for</strong> ρ. y =0atx =0.From symmetry, dy/dx =0atx = L/2.15.5 General solution f(t) = Ae −6t + Be −2t − 3e −4t . (a) No solution, inconsistentboundary conditions; (b) f(t) =2e −6t + e −2t − 3e −4t .15.7 The auxiliary equation has repeated roots <strong>and</strong> the RHS is contained in thecomplementary function. The solution is y(x) =(A+Bx)e −x +2x 2 e −x . y(2) = 5e −2 .15.9 (a) The auxiliary equation has roots 2, 2, −4; (A+Bx)exp2x+C exp(−4x)+2x+1;(b) ∫ multiply through by sinh 2ax <strong>and</strong> note thatcosech 2ax dx =(2a) −1 ln(| tanh ax|); y = B(sinh 2ax) 1/2 (| tanh ax|) A .15.11 Use Laplace trans<strong>for</strong>ms; write s(s + n) −4 as (s + n) −3 − n(s + n) −4 .15.13 L [C(t)] = x 0 (s +8)/[s(s +2)(s + 4)], yieldingC(t) =x 0 [1 + 1 exp(−4t) − 3 exp(−2t)].2 215.15 The characteristic equation is λ 2 − λ − 1=0.u n =[(1+ √ 5) n − (1 − √ 5) n ]/(2 n√ 5).15.17 From u 4 <strong>and</strong> u 5 , P =5,Q= −4. u n =3/2 − 5(−1) n /6+(−2) n /4+2 n /12.15.19 The general solution is A+B2 n/2 exp(i3πn/4)+C2 n/2 exp(i5πn/4). The initial valuesimply that A =1/5,B =( √ 5/10) exp[i(π − φ)] <strong>and</strong> C =( √ 5/10) exp[i(π + φ)].15.21 This is Euler’s equation; setting x =expt produces d 2 z/dt 2 − 2 dz/dt + z =expt,with complementary function (A + Bt)expt <strong>and</strong> particular integral t 2 (exp t)/2;y(x) =x +[x ln x(1 + ln x)]/2.15.23 After multiplication through by x 2 the coefficients are such that this is anexact equation. The resulting first-order equation, in st<strong>and</strong>ard <strong>for</strong>m, needs anintegrating factor (x − 2) 2 /x 2 .15.25 Given the boundary conditions, it is better to work with sinh x <strong>and</strong> sinh(1 − x)than with e ±x ; G(x, ξ) =−[sinh(1 − ξ)sinhx]/ sinh 1 <strong>for</strong> xξ.15.27 Follow the method of subsection 15.2.5, but using general rather than specificfunctions.15.29 G(t, τ) =0<strong>for</strong>tτ. For a unit step input,x(t) =(1+κ 2 ) −1 (1 − e −t cos κt − κ −1 e −t sin κt). Both trans<strong>for</strong>ms are equivalent tos[(s +1) 2 + κ 2 )]¯x =1.529


HIGHER-ORDER ORDINARY DIFFERENTIAL EQUATIONS15.31 Use continuity <strong>and</strong> the step condition on ∂G/∂t at t = t 0 to show thatG(t, t 0 )=α −1 {1 − exp[α(t 0 − t)]} <strong>for</strong> 0 ≤ t 0 ≤ t;x(t) =A(α − a) −1 {a −1 [1 − exp(−at)] − α −1 [1 − exp(−αt)]}.15.33 The LHS of the equation is exact <strong>for</strong> two stages of integration <strong>and</strong> then needsan integrating factor exp x; 2yd 2 y/dx 2 +2ydy/dx+2(dy/dx) 2 ;2ydy/dx+ y 2 =d(y 2 )/dx + y 2 ; y 2 = A exp(−x)+Bx + C − (sin x − cos x)/2.15.35 Follow the method of subsection 15.2.6; u(x) =e −x2 <strong>and</strong> v(x) satisfiesv ′′ +4v =sin 2x, <strong>for</strong> which a particular integral is (−x cos 2x)/4. The general solution isy(x) =[A sin 2x +(B − 1 4 x)cos2x]e−x2 .15.37 The equation is isobaric, with y of weight m, where m + p − 2 = mn; v(x)satisfies x 2 v ′′ + xv ′ = v n .Setx = e t <strong>and</strong> v(x) =u(t), leading to u ′′ = u n withu(0) = 0,u ′ (0) = λ. Multiply both sides by u ′ to make the equation exact.530


16Series solutions of ordinarydifferential equationsIn the previous chapter the solution of both homogeneous <strong>and</strong> non-homogeneouslinear ordinary differential equations (ODEs) of order ≥ 2 was discussed. In particularwe developed methods <strong>for</strong> solving some equations in which the coefficientswere not constant but functions of the independent variable x. In each case wewere able to write the solutions to such equations in terms of elementary functions,or as integrals. In general, however, the solutions of equations with variablecoefficients cannot be written in this way, <strong>and</strong> we must consider alternativeapproaches.In this chapter we discuss a method <strong>for</strong> obtaining solutions to linear ODEsin the <strong>for</strong>m of convergent series. Such series can be evaluated numerically, <strong>and</strong>those occurring most commonly are named <strong>and</strong> tabulated. There is in fact nodistinct borderline between this <strong>and</strong> the previous chapter, since solutions in termsof elementary functions may equally well be written as convergent series (i.e. therelevant Taylor series). Indeed, it is partly because some series occur so frequentlythat they are given special names such as sin x, cosx or exp x.Since we shall be concerned principally with second-order linear ODEs in thischapter, we begin with a discussion of these equations, <strong>and</strong> obtain some generalresults that will prove useful when we come to discuss series solutions.16.1 Second-order linear ordinary differential equationsAny homogeneous second-order linear ODE can be written in the <strong>for</strong>my ′′ + p(x)y ′ + q(x)y =0, (16.1)where y ′ = dy/dx <strong>and</strong> p(x) <strong>and</strong>q(x) are given functions of x. From the previouschapter, we recall that the most general <strong>for</strong>m of the solution to (16.1) isy(x) =c 1 y 1 (x)+c 2 y 2 (x), (16.2)531


SERIES SOLUTIONS OF ORDINARY DIFFERENTIAL EQUATIONSwhere y 1 (x) <strong>and</strong>y 2 (x) arelinearly independent solutions of (16.1), <strong>and</strong> c 1 <strong>and</strong> c 2are constants that are fixed by the boundary conditions (if supplied).A full discussion of the linear independence of sets of functions was givenat the beginning of the previous chapter, but <strong>for</strong> just two functions y 1 <strong>and</strong> y 2to be linearly independent we simply require that y 2 is not a multiple of y 1 .Equivalently, y 1 <strong>and</strong> y 2 must be such that the equationc 1 y 1 (x)+c 2 y 2 (x) =0is only satisfied <strong>for</strong> c 1 = c 2 = 0. There<strong>for</strong>e the linear independence of y 1 (x) <strong>and</strong>y 2 (x) can usually be deduced by inspection but in any case can always be verifiedby the evaluation of the Wronskian of the two solutions,W (x) =∣ y 1 y 2y 1 ′ y 2′ ∣ = y 1y 2 ′ − y 2 y 1. ′ (16.3)If W (x) ≠ 0 anywhere in a given interval then y 1 <strong>and</strong> y 2 are linearly independentin that interval.An alternative expression <strong>for</strong> W (x), of which we will make use later, may bederived by differentiating (16.3) with respect to x to giveW ′ = y 1 y 2 ′′ + y 1y ′ 2 ′ − y 2 y 1 ′′ − y 2y ′ 1 ′ = y 1 y 2 ′′ − y 1y ′′2 .Since both y 1 <strong>and</strong> y 2 satisfy (16.1), we may substitute <strong>for</strong> y 1 ′′ <strong>and</strong> y′′ 2 to obtainW ′ = −y 1 (py 2 ′ + qy 2 )+(py 1 ′ + qy 1 )y 2 = −p(y 1 y 2 ′ − y 1y ′ 2 )=−pW .Integrating, we find{ ∫ x}W (x) =C exp − p(u) du , (16.4)where C is a constant. We note further that in the special case p(x) ≡ 0weobtainW =constant.◮The functions y 1 =sinx <strong>and</strong> y 2 =cosx are both solutions of the equation y ′′ + y =0. Evaluate the Wronskian of these two solutions, <strong>and</strong> hence show that they are linearlyindependent.The Wronskian of y 1 <strong>and</strong> y 2 is given byW = y 1 y 2 ′ − y 2 y 1 ′ = − sin 2 x − cos 2 x = −1.Since W ≠ 0 the two solutions are linearly independent. We also note that y ′′ + y =0isa special case of (16.1) with p(x) = 0. We there<strong>for</strong>e expect, from (16.4), that W will be aconstant, as is indeed the case. ◭From the previous chapter we recall that, once we have obtained the generalsolution to the homogeneous second-order ODE (16.1) in the <strong>for</strong>m (16.2), thegeneral solution to the inhomogeneous equationy ′′ + p(x)y ′ + q(x)y = f(x) (16.5)532


16.1 SECOND-ORDER LINEAR ORDINARY DIFFERENTIAL EQUATIONScan be written as the sum of the solution to the homogeneous equation y c (x)(the complementary function) <strong>and</strong> any function y p (x) (the particular integral) thatsatisfies (16.5) <strong>and</strong> is linearly independent of y c (x). We have there<strong>for</strong>ey(x) =c 1 y 1 (x)+c 2 y 2 (x)+y p (x). (16.6)General methods <strong>for</strong> obtaining y p , that are applicable to equations with variablecoefficients, such as the variation of parameters or Green’s functions, were discussedin the previous chapter. An alternative description of the Green’s functionmethod <strong>for</strong> solving inhomogeneous equations is given in the next chapter. For thepresent, however, we will restrict our attention to the solutions of homogeneousODEs in the <strong>for</strong>m of convergent series.16.1.1 Ordinary <strong>and</strong> singular points of an ODESo far we have implicitly assumed that y(x) isareal function of a real variablex. However, this is not always the case, <strong>and</strong> in the remainder of this chapter webroaden our discussion by generalising to a complex function y(z) ofacomplexvariable z.Let us there<strong>for</strong>e consider the second-order linear homogeneous ODEy ′′ + p(z)y ′ + q(z) =0, (16.7)where now y ′ = dy/dz; this is a straight<strong>for</strong>ward generalisation of (16.1). A fulldiscussion of complex functions <strong>and</strong> differentiation with respect to a complexvariable z is given in chapter 24, but <strong>for</strong> the purposes of the present chapter weneed not concern ourselves with many of the subtleties that exist. In particular,we may treat differentiation with respect to z in a way analogous to ordinarydifferentiation with respect to a real variable x.In (16.7), if, at some point z = z 0 , the functions p(z) <strong>and</strong>q(z) are finite <strong>and</strong> canbe expressed as complex power series (see section 4.5), i.e.∞∑∞∑p(z) = p n (z − z 0 ) n , q(z) = q n (z − z 0 ) n ,n=0then p(z) <strong>and</strong>q(z) are said to be analytic at z = z 0 , <strong>and</strong> this point is called anordinary point of the ODE. If, however, p(z) orq(z), or both, diverge at z = z 0then it is called a singular point of the ODE.Even if an ODE is singular at a given point z = z 0 , it may still possess anon-singular (finite) solution at that point. In fact the necessary <strong>and</strong> sufficientcondition § <strong>for</strong> such a solution to exist is that (z − z 0 )p(z) <strong>and</strong>(z − z 0 ) 2 q(z) areboth analytic at z = z 0 . Singular points that have this property are called regularn=0§ See, <strong>for</strong> example, H. Jeffreys <strong>and</strong> B. S. Jeffreys, <strong>Methods</strong> of <strong>Mathematical</strong> <strong>Physics</strong>, 3rdedn(Cambridge:Cambridge University Press, 1966), p. 479.533


SERIES SOLUTIONS OF ORDINARY DIFFERENTIAL EQUATIONSsingular points, whereas any singular point not satisfying both these criteria istermed an irregular or essential singularity.◮Legendre’s equation has the <strong>for</strong>m(1 − z 2 )y ′′ − 2zy ′ + l(l +1)y =0, (16.8)where l is a constant. Show that z =0is an ordinary point <strong>and</strong> z = ±1 are regular singularpoints of this equation.Firstly, divide through by 1 − z 2 to put the equation into our st<strong>and</strong>ard <strong>for</strong>m (16.7):y ′′ −2z l(l +1)1 − z 2 y′ +1 − z y =0.2Comparing this with (16.7), we identify p(z) <strong>and</strong>q(z) asp(z) =−2z1 − z = −2zl(l +1) l(l +1), q(z) = =2 (1 + z)(1 − z) 1 − z 2 (1 + z)(1 − z) .By inspection, p(z) <strong>and</strong>q(z) are analytic at z = 0, which is there<strong>for</strong>e an ordinary point,but both diverge <strong>for</strong> z = ±1, which are thus singular points. However, at z =1weseethat both (z − 1)p(z) <strong>and</strong>(z − 1) 2 q(z) are analytic <strong>and</strong> hence z = 1 is a regular singularpoint. Similarly, at z = −1 both (z +1)p(z) <strong>and</strong>(z +1) 2 q(z) are analytic, <strong>and</strong> it too is aregular singular point. ◭So far we have assumed that z 0 is finite. However, we may sometimes wish todetermine the nature of the point |z| →∞. This may be achieved straight<strong>for</strong>wardlyby substituting w =1/z into the equation <strong>and</strong> investigating the behaviour atw =0.◮Show that Legendre’s equation has a regular singularity at |z| →∞.Letting w =1/z, the derivatives with respect to z becomedydz = dy dwdw dz = − 1 dy dy=z 2 −w2dw dw ,d 2 ydz = dw ( ) (d dy= −w 2 −2w dy ) (2 dz dw dzdw − d2 yw2 = w 3 2 dy )dw 2 dw + w d2 y.dw 2If we substitute these derivatives into Legendre’s equation (16.8) we obtain(1 − 1 ) (w 3 2 dy )w 2 dw + w d2 y+2 1 dydw 2 w2 + l(l +1)y =0,w dwwhich simplifies to givew 2 (w 2 − 1) d2 y dy+2w3 + l(l +1)y =0.dw2 dwDividing through by w 2 (w 2 − 1) to put the equation into st<strong>and</strong>ard <strong>for</strong>m, <strong>and</strong> comparingwith (16.7), we identify p(w) <strong>and</strong>q(w) asp(w) =2wl(l +1), q(w) =w 2 − 1 w 2 (w 2 − 1) .At w =0,p(w) is analytic but q(w) diverges, <strong>and</strong> so the point |z| →∞is a singular pointof Legendre’s equation. However, since wp <strong>and</strong> w 2 q are both analytic at w =0,|z| →∞is a regular singular point. ◭534


16.2 SERIES SOLUTIONS ABOUT AN ORDINARY POINTEquation Regular Essentialsingularities singularitiesHypergeometricz(1 − z)y ′′ +[c − (a + b +1)z]y ′ − aby =0 0, 1, ∞ —Legendre(1 − z 2 )y ′′ − 2zy ′ + l(l +1)y =0 −1, 1, ∞ —Associated Legendre](1 − z 2 )y ′′ − 2zy ′ +[l(l +1)− m2y =0 −1, 1, ∞ —1 − z 2Chebyshev(1 − z 2 )y ′′ − zy ′ + ν 2 y =0 −1, 1, ∞ —Confluent hypergeometriczy ′′ +(c − z)y ′ − ay =0 0 ∞Besselz 2 y ′′ + zy ′ +(z 2 − ν 2 )y =0 0 ∞Laguerrezy ′′ +(1− z)y ′ + νy =0 0 ∞Associated Laguerrezy ′′ +(m +1− z)y ′ +(ν − m)y =0 0 ∞Hermitey ′′ − 2zy ′ +2νy =0 — ∞Simple harmonic oscillatory ′′ + ω 2 y =0 — ∞Table 16.1 Important second-order linear ODEs in the physical sciences <strong>and</strong>engineering.Table 16.1 lists the singular points of several second-order linear ODEs thatplay important roles in the analysis of many problems in physics <strong>and</strong> engineering.A full discussion of the solutions to each of the equations in table 16.1 <strong>and</strong> theirproperties is left until chapter 18. We now discuss the general methods by whichseries solutions may be obtained.16.2 Series solutions about an ordinary pointIf z = z 0 is an ordinary point of (16.7) then it may be shown that every solutiony(z) of the equation is also analytic at z = z 0 . From now on we will take z 0 as theorigin, i.e. z 0 = 0. If this is not already the case, then a substitution Z = z − z 0will make it so. Since every solution is analytic, y(z) can be represented by a535


SERIES SOLUTIONS OF ORDINARY DIFFERENTIAL EQUATIONSpower series of the <strong>for</strong>m (see section 24.11)∞∑y(z) = a n z n . (16.9)n=0Moreover, it may be shown that such a power series converges <strong>for</strong> |z|


16.2 SERIES SOLUTIONS ABOUT AN ORDINARY POINTagivena 0 . Alternatively, starting with a 1 , we obtain the odd coefficients a 3 ,a 5 ,etc.Twoindependent solutions of the ODE can be obtained by setting either a 0 =0ora 1 =0.Firstly, if we set a 1 = 0 <strong>and</strong> choose a 0 = 1 then we obtain the solution∞y 1 (z) =1− z22! + z44! − ···= ∑ (−1) n(2n)! z2n .n=0Secondly, if we set a 0 = 0 <strong>and</strong> choose a 1 = 1 then we obtain a second, independent, solution∞y 2 (z) =z − z33! + z55! − ···= ∑ (−1) n(2n +1)! z2n+1 .n=0Recognising these two series as cos z <strong>and</strong> sin z, we can write the general solution asy(z) =c 1 cos z + c 2 sin z,where c 1 <strong>and</strong> c 2 are arbitrary constants that are fixed by boundary conditions (if supplied).We note that both solutions converge <strong>for</strong> all z, as might be expected since the ODEpossesses no singular points (except |z| →∞). ◭Solving the above example was quite straight<strong>for</strong>ward <strong>and</strong> the resulting serieswere easily recognised <strong>and</strong> written in closed <strong>for</strong>m (i.e. in terms of elementaryfunctions); this is not usually the case. Another simplifying feature of the previousexample was that we obtained a two-term recurrence relation relating a n+2 <strong>and</strong>a n , so that the odd- <strong>and</strong> even-numbered coefficients were independent of oneanother. In general, the recurrence relation expresses a n in terms of any numberof the previous a r (0 ≤ r ≤ n − 1).◮Find the series solutions, about z =0,ofy ′′ −2(1 − z) 2 y =0.By inspection, z = 0 is an ordinary point, <strong>and</strong> there<strong>for</strong>e we may find two independentsolutions by substituting y = ∑ ∞n=0 a nz n . Using (16.10) <strong>and</strong> (16.11), <strong>and</strong> multiplying throughby (1 − z) 2 , we find∞∑∞∑(1 − 2z + z 2 ) n(n − 1)a n z n−2 − 2 a n z n =0,n=0n=0n=0which leads to∞∑∞∑∞∑∞∑n(n − 1)a n z n−2 − 2 n(n − 1)a n z n−1 + n(n − 1)a n z n − 2 a n z n =0.In order to write all these series in terms of the coefficients of z n , we must shift thesummation index in the first two sums, obtaining∞∑∞∑∞∑(n +2)(n +1)a n+2 z n − 2 (n +1)na n+1 z n + (n 2 − n − 2)a n z n =0,n=0which can be written as∞∑(n +1)[(n +2)a n+2 − 2na n+1 +(n − 2)a n ]z n =0.n=0n=0537n=0n=0n=0n=0


SERIES SOLUTIONS OF ORDINARY DIFFERENTIAL EQUATIONSBy dem<strong>and</strong>ing that the coefficients of each power of z vanish separately, we obtain thethree-term recurrence relation(n +2)a n+2 − 2na n+1 +(n − 2)a n =0 <strong>for</strong>n ≥ 0,which determines a n <strong>for</strong> n ≥ 2intermsofa 0 <strong>and</strong> a 1 . Three-term (or more) recurrencerelations are a nuisance <strong>and</strong>, in general, can be difficult to solve. This particular recurrencerelation, however, has two straight<strong>for</strong>ward solutions. One solution is a n = a 0 <strong>for</strong> all n, inwhich case (choosing a 0 = 1) we findy 1 (z) =1+z + z 2 + z 3 + ···= 11 − z .The other solution to the recurrence relation is a 1 = −2a 0 , a 2 = a 0 <strong>and</strong> a n =0<strong>for</strong>n>2,so that (again choosing a 0 =1)weobtainapolynomial solution to the ODE:y 2 (z) =1− 2z + z 2 =(1− z) 2 .The linear independence of y 1 <strong>and</strong> y 2 is obvious but can be checked by computing theWronskianW = y 1 y 2 ′ − y 1y ′ 2 = 11 − z [−2(1 − z)] − 1(1 − z) (1 − 2 z)2 = −3.Since W ≠ 0, the two solutions y 1 <strong>and</strong> y 2 are indeed linearly independent. The generalsolution of the ODE is there<strong>for</strong>ey(z) = c 11 − z + c 2(1 − z) 2 .We observe that y 1 (<strong>and</strong> hence the general solution) is singular at z =1,whichisthesingular point of the ODE nearest to z = 0, but the polynomial solution, y 2 , is valid <strong>for</strong>all finite z. ◭The above example illustrates the possibility that, in some cases, we may findthat the recurrence relation leads to a n =0<strong>for</strong>n>N, <strong>for</strong> one or both of thetwo solutions; we then obtain a polynomial solution to the equation. Polynomialsolutions are discussed more fully in section 16.5, but one obvious property ofsuch solutions is that they converge <strong>for</strong> all finite z. By contrast, as mentionedabove, <strong>for</strong> solutions in the <strong>for</strong>m of an infinite series the circle of convergenceextends only as far as the singular point nearest to that about which the solutionis being obtained.16.3 Series solutions about a regular singular pointFrom table 16.1 we see that several of the most important second-order linearODEs in physics <strong>and</strong> engineering have regular singular points in the finite complexplane. We must extend our discussion, there<strong>for</strong>e, to obtaining series solutions toODEs about such points. In what follows we assume that the regular singularpoint about which the solution is required is at z = 0, since, as we have seen, ifthis is not already the case then a substitution of the <strong>for</strong>m Z = z − z 0 will makeit so.If z = 0 is a regular singular point of the equationy ′′ + p(z)y ′ + q(z)y =0538


16.3 SERIES SOLUTIONS ABOUT A REGULAR SINGULAR POINTthen at least one of p(z) <strong>and</strong>q(z) is not analytic at z = 0, <strong>and</strong> in general weshould not expect to find a power series solution of the <strong>for</strong>m (16.9). We mustthere<strong>for</strong>e extend the method to include a more general <strong>for</strong>m <strong>for</strong> the solution. Infact, it may be shown (Fuch’s theorem) that there exists at least one solution tothe above equation, of the <strong>for</strong>my = z σ∞ ∑n=0a n z n , (16.12)where the exponent σ is a number that may be real or complex <strong>and</strong> where a 0 ≠0(since, if it were otherwise, σ could be redefined as σ +1orσ +2or··· so as tomake a 0 ≠ 0). Such a series is called a generalised power series or Frobenius series.As in the case of a simple power series solution, the radius of convergence of theFrobenius series is, in general, equal to the distance to the nearest singularity ofthe ODE.Since z = 0 is a regular singularity of the ODE, it follows that zp(z) <strong>and</strong>z 2 q(z)are analytic at z = 0, so that we may write∞∑zp(z) ≡ s(z) = s n z n ,z 2 q(z) ≡ t(z) =n=0∞∑t n z n ,where we have defined the analytic functions s(z) <strong>and</strong>t(z) <strong>for</strong> later convenience.The original ODE there<strong>for</strong>e becomesy ′′ + s(z)z y′ + t(z)z 2 y =0.Let us substitute the Frobenius series (16.12) into this equation. The derivativesof (16.12) with respect to z are given by∞∑y ′ = (n + σ)a n z n+σ−1 , (16.13)n=0y ′′ =n=0n=0∞∑(n + σ)(n + σ − 1)a n z n+σ−2 , (16.14)n=0<strong>and</strong> we obtain∞∑∞∑∞∑(n + σ)(n + σ − 1)a n z n+σ−2 + s(z) (n + σ)a n z n+σ−2 + t(z) a n z n+σ−2 =0.n=0Dividing this equation through by z σ−2 , we find∞∑[(n + σ)(n + σ − 1) + s(z)(n + σ)+t(z)] a n z n =0. (16.15)n=0539n=0


SERIES SOLUTIONS OF ORDINARY DIFFERENTIAL EQUATIONSSetting z = 0, all terms in the sum with n>0 vanish, implying that[σ(σ − 1) + s(0)σ + t(0)]a 0 =0,which, since we require a 0 ≠ 0, yields the indicial equationσ(σ − 1) + s(0)σ + t(0) = 0. (16.16)This equation is a quadratic in σ <strong>and</strong> in general has two roots, the nature ofwhich determines the <strong>for</strong>ms of possible series solutions.The two roots of the indicial equation, σ 1 <strong>and</strong> σ 2 , are called the indices ofthe regular singular point. By substituting each of these roots into (16.15) inturn <strong>and</strong> requiring that the coefficients of each power of z vanish separately, weobtain a recurrence relation (<strong>for</strong> each root) expressing each a n as a function ofthe previous a r (0 ≤ r ≤ n − 1). We will see that the larger root of the indicialequation always yields a solution to the ODE in the <strong>for</strong>m of a Frobenius series(16.12). The <strong>for</strong>m of the second solution depends, however, on the relationshipbetween the two indices σ 1 <strong>and</strong> σ 2 . There are three possible general cases: (i)distinct roots not differing by an integer; (ii) repeated roots; (iii) distinct rootsdiffering by an integer (not equal to zero). Below, we discuss each of these in turn.Be<strong>for</strong>e continuing, however, we note that, as was the case <strong>for</strong> solutions inthe <strong>for</strong>m of a simple power series, it is always worth investigating whether aFrobenius series found as a solution to a problem is summable in closed <strong>for</strong>mor expressible in terms of known functions. We illustrate this point below, butthe reader should avoid gaining the impression that this is always so or that, ifone worked hard enough, a closed-<strong>for</strong>m solution could always be found withoutusing the series method. As mentioned earlier, this is not the case, <strong>and</strong> very oftenan infinite series solution is the best one can do.16.3.1 Distinct roots not differing by an integerIf the roots of the indicial equation, σ 1 <strong>and</strong> σ 2 , differ by an amount that is notan integer then the recurrence relations corresponding to each root lead to twolinearly independent solutions of the ODE:y 1 (z) =z σ1∞ ∑n=0∑∞a n z n , y 2 (z) =z σ2 b n z n ,with both solutions taking the <strong>for</strong>m of a Frobenius series. The linear independenceof these two solutions follows from the fact that y 2 /y 1 is not a constant sinceσ 1 − σ 2 is not an integer. Because y 1 <strong>and</strong> y 2 are linearly independent, we may usethem to construct the general solution y = c 1 y 1 + c 2 y 2 .We also note that this case includes complex conjugate roots where σ 2 = σ ∗ 1 ,since σ 1 − σ 2 = σ 1 − σ ∗ 1 =2i Im σ 1 cannot be equal to a real integer.540n=0


16.3 SERIES SOLUTIONS ABOUT A REGULAR SINGULAR POINT◮Find the power series solutions about z =0of4zy ′′ +2y ′ + y =0.Dividing through by 4z to put the equation into st<strong>and</strong>ard <strong>for</strong>m, we obtainy ′′ + 1 2z y′ + 1 y =0, (16.17)4z<strong>and</strong> on comparing with (16.7) we identify p(z) =1/(2z) <strong>and</strong>q(z) =1/(4z). Clearly z =0is a singular point of (16.17), but since zp(z) =1/2 <strong>and</strong>z 2 q(z) =z/4 are finite there, itis a regular singular point. We there<strong>for</strong>e substitute the Frobenius series y = z ∑ σ ∞n=0 a nz ninto (16.17). Using (16.13) <strong>and</strong> (16.14), we obtain∞∑(n + σ)(n + σ − 1)a n z n+σ−2 + 1 ∞∑(n + σ)a n z n+σ−1 + 1 ∞∑a n z n+σ =0,2z4zn=0n=0which, on dividing through by z σ−2 ,gives∞∑ [(n + σ)(n + σ − 1) +1(n + σ)+ 1 z] a2 4 n z n =0. (16.18)n=0If we set z = 0 then all terms in the sum with n>0 vanish, <strong>and</strong> we obtain the indicialequationσ(σ − 1) + 1 σ =0, 2which has roots σ =1/2 <strong>and</strong>σ = 0. Since these roots do not differ by an integer, weexpect to find two independent solutions to (16.17), in the <strong>for</strong>m of Frobenius series.Dem<strong>and</strong>ing that the coefficients of z n vanish separately in (16.18), we obtain therecurrence relation(n + σ)(n + σ − 1)a n + 1 (n + σ)a 2 n + 1 a 4 n−1 =0. (16.19)If we choose the larger root, σ =1/2, of the indicial equation then (16.19) becomes(4n 2 +2n)a n + a n−1 =0 ⇒ a n = −a n−12n(2n +1) .Setting a 0 = 1, we find a n =(−1) n /(2n + 1)!, <strong>and</strong> so the solution to (16.17) is given byy 1 (z) = √ z∞∑n=0= √ z − (√ z) 33!(−1) n(2n +1)! zn+ (√ z) 55!− ···=sin √ z.To obtain the second solution we set σ = 0 (the smaller root of the indicial equation) in(16.19), which gives(4n 2 − 2n)a n + a n−1 =0 ⇒ a n = − a n−12n(2n − 1) .Setting a 0 =1nowgivesa n =(−1) n /(2n)!, <strong>and</strong> so the second (independent) solution to(16.17) isy 2 (z) =∞∑n=0n=0(−1) n(2n)! zn =1− (√ z) 2+ (√ 4) 4− ···=cos √ z.2! 4!541


SERIES SOLUTIONS OF ORDINARY DIFFERENTIAL EQUATIONSWe may check that y 1 (z) <strong>and</strong>y 2 (z) are indeed linearly independent by computing theWronskian as follows:W = y 1 y 2 ′ − y 2 y 1′=sin √ (z − 12 √ z sin √ )z − cos √ ( 1z2 √ z cos √ )z= − 12 √ ( √ sin2z +cos √ 2 z ) = − 1z2 √ z ≠0.Since W ≠0,thesolutionsy 1 (z) <strong>and</strong>y 2 (z) are linearly independent. Hence, the generalsolution to (16.17) is given byy(z) =c 1 sin √ z + c 2 cos √ z. ◭16.3.2 Repeated root of the indicial equationIf the indicial equation has a repeated root, so that σ 1 = σ 2 = σ, then obviouslyonly one solution in the <strong>for</strong>m of a Frobenius series (16.12) may be found asdescribed above, i.e.y 1 (z) =z σ∞ ∑n=0a n z n .<strong>Methods</strong> <strong>for</strong> obtaining a second, linearly independent, solution are discussed insection 16.4.16.3.3 Distinct roots differing by an integerWhatever the roots of the indicial equation, the recurrence relation correspondingto the larger of the two always leads to a solution of the ODE. However, if theroots of the indicial equation differ by an integer then the recurrence relationcorresponding to the smaller root may or may not lead to a second linearlyindependent solution, depending on the ODE under consideration. Note that <strong>for</strong>complex roots of the indicial equation, the ‘larger’ root is taken to be the onewith the larger real part.◮Find the power series solutions about z =0ofz(z − 1)y ′′ +3zy ′ + y =0. (16.20)Dividing through by z(z − 1) to put the equation into st<strong>and</strong>ard <strong>for</strong>m, we obtainy ′′ + 3 1(z − 1) y′ + y =0, (16.21)z(z − 1)<strong>and</strong> on comparing with (16.7) we identify p(z) =3/(z − 1) <strong>and</strong> q(z) =1/[z(z − 1)]. Weimmediately see that z = 0 is a singular point of (16.21), but since zp(z) =3z/(z − 1) <strong>and</strong>z 2 q(z) =z/(z−1) are finite there, it is a regular singular point <strong>and</strong> we expect to find at least542


16.3 SERIES SOLUTIONS ABOUT A REGULAR SINGULAR POINTone solution in the <strong>for</strong>m of a Frobenius series. We there<strong>for</strong>e substitute y = z ∑ σ ∞n=0 a nz ninto (16.21) <strong>and</strong>, using (16.13) <strong>and</strong> (16.14), we obtain∞∑(n + σ)(n + σ − 1)a n z n+σ−2 + 3 ∞∑(n + σ)a n z n+σ−1z − 1n=0which, on dividing through by z σ−2 ,gives∞∑n=0n=0[(n + σ)(n + σ − 1) + 3z (n + σ)+zz − 1+1z(z − 1)∞∑a n z n+σ =0,n=0]a n z n =0.z − 1Although we could use this expression to find the indicial equation <strong>and</strong> recurrence relations,the working is simpler if we now multiply through by z − 1togive∞∑[(z − 1)(n + σ)(n + σ − 1) + 3z(n + σ)+z] a n z n =0. (16.22)n=0If we set z = 0 then all terms in the sum with the exponent of z greater than zero vanish,<strong>and</strong> we obtain the indicial equationσ(σ − 1) = 0,which has the roots σ =1<strong>and</strong>σ = 0. Since the roots differ by an integer (unity), it may notbe possible to find two linearly independent solutions of (16.21) in the <strong>for</strong>m of Frobeniusseries. We are guaranteed, however, to find one such solution corresponding to the largerroot, σ =1.Dem<strong>and</strong>ing that the coefficients of z n vanish separately in (16.22), we obtain therecurrence relation(n − 1+σ)(n − 2+σ)a n−1 − (n + σ)(n + σ − 1)a n +3(n − 1+σ)a n−1 + a n−1 =0,which can be simplified to give(n + σ − 1)a n =(n + σ)a n−1 . (16.23)On substituting σ = 1 into this expression, we obtain( ) n +1a n = a n−1 ,n<strong>and</strong>onsettinga 0 = 1 we find a n = n + 1; so one solution to (16.21) is given by∞∑y 1 (z) =z (n +1)z n = z(1+2z +3z 2 + ···)n=0z=(1 − z) . (16.24)2If we attempt to find a second solution (corresponding to the smaller root of the indicialequation) by setting σ = 0 in (16.23), we find( n)a n = a n−1 .n − 1But we require a 0 ≠0,soa 1 is <strong>for</strong>mally infinite <strong>and</strong> the method fails. We discuss how tofind a second linearly independent solution in the next section. ◭One particular case is worth mentioning. If the point about which the solution543


SERIES SOLUTIONS OF ORDINARY DIFFERENTIAL EQUATIONSis required, i.e. z = 0, is in fact an ordinary point of the ODE rather than aregular singular point, then substitution of the Frobenius series (16.12) leads toan indicial equation with roots σ = 0 <strong>and</strong> σ = 1. Although these roots differ byan integer (unity), the recurrence relations corresponding to the two roots yieldtwo linearly independent power series solutions (one <strong>for</strong> each root), as expectedfrom section 16.2.16.4 Obtaining a second solutionWhilst attempting to construct solutions to an ODE in the <strong>for</strong>m of Frobeniusseries about a regular singular point, we found in the previous section that whenthe indicial equation has a repeated root, or roots differing by an integer, we can(in general) find only one solution of this <strong>for</strong>m. In order to construct the generalsolution to the ODE, however, we require two linearly independent solutions y 1<strong>and</strong> y 2 . We now consider several methods <strong>for</strong> obtaining a second solution in thiscase.16.4.1 The Wronskian methodIf y 1 <strong>and</strong> y 2 are two linearly independent solutions of the st<strong>and</strong>ard equationy ′′ + p(z)y ′ + q(z)y =0then the Wronskian of these two solutions is given by W (z) =y 1 y 2 ′ − y 2y 1 ′ .Dividing the Wronskian by y1 2 we obtain[ ( )]Wy12 = y′ 2− y′ 1y 1 y12 y 2 = y′ 2 d 1+ y 2 = d ( )y2,y 1 dz y 1 dz y 1which integrates to give∫ zW (u)y 2 (z) =y 1 (z)y1 2 du.(u)Now using the alternative expression <strong>for</strong> W (z) given in (16.4) with C =1(sincewe are not concerned with this normalising factor), we find∫ z{ ∫1u}y 2 (z) =y 1 (z)y1 2(u) exp − p(v) dv du. (16.25)Hence, given y 1 , we can in principle compute y 2 . Note that the lower limits ofintegration have been omitted. If constant lower limits are included then theymerely lead to a constant times the first solution.◮Find a second solution to (16.21) using the Wronskian method.For the ODE (16.21) we have p(z) =3/(z − 1), <strong>and</strong> from (16.24) we see that one solution544


16.4 OBTAINING A SECOND SOLUTIONto (16.21) is y 1 = z/(1 − z) 2 . Substituting <strong>for</strong> p <strong>and</strong> y 1 in (16.25) we have∫z z((1 − u) 4 ∫ u) 3y 2 (z) =exp −(1 − z) 2 u 2v − 1 dv du∫z z(1 − u) 4=(1 − z) 2 u∫ 2z zu − 1=du(1 − z) 2 u 2=z(1 − z) 2 (ln z + 1 z).exp [−3ln(u − 1)] duBy calculating the Wronskian of y 1 <strong>and</strong> y 2 it is easily shown that, as expected, the twosolutions are linearly independent. In fact, as the Wronskian has already been evaluatedas W (u) =exp[−3ln(u − 1)], i.e. W (z) =(z − 1) −3 , no calculation is needed. ◭An alternative (but equivalent) method of finding a second solution is simply toassume that the second solution has the <strong>for</strong>m y 2 (z) =u(z)y 1 (z) <strong>for</strong> some functionu(z) to be determined (this method was discussed more fully in subsection 15.2.3).From (16.25), we see that the second solution derived from the Wronskian isindeed of this <strong>for</strong>m. Substituting y 2 (z) =u(z)y 1 (z) into the ODE leads to afirst-order ODE in which u ′ is the dependent variable; this may then be solved.16.4.2 The derivative methodThe derivative method of finding a second solution begins with the derivation ofa recurrence relation <strong>for</strong> the coefficients a n in a Frobenius series solution, as in theprevious section. However, rather than putting σ = σ 1 in this recurrence relationto evaluate the first series solution, we now keep σ as a variable parameter. Thismeans that the computed a n are functions of σ <strong>and</strong> the computed solution is nowa function of z <strong>and</strong> σ:y(z,σ) =z σ∞ ∑n=0a n (σ)z n . (16.26)Of course, if we put σ = σ 1 in this, we obtain immediately the first series solution,but <strong>for</strong> the moment we leave σ as a parameter.For brevity let us denote the differential operator on the LHS of our st<strong>and</strong>ardODE (16.7) by L, sothatL = d2dz 2 + p(z) d dz + q(z),<strong>and</strong> examine the effect of L on the series y(z,σ) in (16.26). It is clear that theseries Ly(z,σ) will contain only a term in z σ , since the recurrence relation definingthe a n (σ) is such that these coefficients vanish <strong>for</strong> higher powers of z. Butthecoefficient of z σ is simply the LHS of the indicial equation. There<strong>for</strong>e, if the roots545


SERIES SOLUTIONS OF ORDINARY DIFFERENTIAL EQUATIONSof the indicial equation are σ = σ 1 <strong>and</strong> σ = σ 2 then it follows thatLy(z,σ) =a 0 (σ − σ 1 )(σ − σ 2 )z σ . (16.27)There<strong>for</strong>e, as in the previous section, we see that <strong>for</strong> y(z,σ) tobeasolutionofthe ODE Ly =0,σ must equal σ 1 or σ 2 . For simplicity we shall set a 0 = 1 in thefollowing discussion.Let us first consider the case in which the two roots of the indicial equationare equal, i.e. σ 2 = σ 1 . From (16.27) we then haveLy(z,σ) =(σ − σ 1 ) 2 z σ .Differentiating this equation with respect to σ we obtain∂∂σ [Ly(z,σ)] =(σ − σ 1) 2 z σ ln z +2(σ − σ 1 )z σ ,which equals zero if σ = σ 1 . But since ∂/∂σ <strong>and</strong> L are operators that differentiatewith respect to different variables, we can reverse their order, implying that[ ] ∂L∂σ y(z,σ) =0 atσ = σ 1 .Hence, the function in square brackets, evaluated at σ = σ 1 <strong>and</strong> denoted by[ ] ∂∂σ y(z,σ) , (16.28)σ=σ 1is also a solution of the original ODE Ly = 0, <strong>and</strong> is in fact the second linearlyindependent solution that we were looking <strong>for</strong>.The case in which the roots of the indicial equation differ by an integer isslightly more complicated but can be treated in a similar way. In (16.27), since Ldifferentiates with respect to z we may multiply (16.27) by any function of σ, sayσ − σ 2 , <strong>and</strong> take this function inside the operator L on the LHS to obtainL [(σ − σ 2 )y(z,σ)] =(σ − σ 1 )(σ − σ 2 ) 2 z σ . (16.29)There<strong>for</strong>e the function[(σ − σ 2 )y(z,σ)] σ=σ2is also a solution of the ODE Ly = 0. However, it can be proved § that thisfunction is a simple multiple of the first solution y(z,σ 1 ), showing that it is notlinearly independent <strong>and</strong> that we must find another solution. To do this wedifferentiate (16.29) with respect to σ <strong>and</strong> find∂∂σ {L [(σ − σ 2)y(z,σ)]} =(σ − σ 2 ) 2 z σ +2(σ − σ 1 )(σ − σ 2 )z σ+(σ − σ 1 )(σ − σ 2 ) 2 z σ ln z,§ For a fuller discussion see, <strong>for</strong> example, K. F. Riley, <strong>Mathematical</strong> <strong>Methods</strong> <strong>for</strong> the Physical Sciences(Cambridge: Cambridge University Press, 1974), pp. 158–9.546


16.4 OBTAINING A SECOND SOLUTIONwhich is equal to zero if σ = σ 2 . As previously, since ∂/∂σ <strong>and</strong> L are operatorsthat differentiate with respect to different variables, we can reverse their order toobtain{ }∂L∂σ [(σ − σ 2)y(z,σ)] =0 atσ = σ 2 ,<strong>and</strong> so the function{ }∂∂σ [(σ − σ 2)y(z,σ)](16.30)σ=σ 2is also a solution of the original ODE Ly = 0, <strong>and</strong> is in fact the second linearlyindependent solution.◮Find a second solution to (16.21) using the derivative method.From (16.23) the recurrence relation (with σ as a parameter) is given by(n + σ − 1)a n =(n + σ)a n−1 .Setting a 0 = 1 we find that the coefficients have the particularly simple <strong>for</strong>m a n (σ) =(σ + n)/σ. We there<strong>for</strong>e consider the function∞∑y(z, σ) =z σ a n (σ)z n = z σ σ + nσ zn .n=0n=0The smaller root of the indicial equation <strong>for</strong> (16.21) is σ 2 = 0, <strong>and</strong> so from (16.30) asecond, linearly independent, solution to the ODE is given by{ { [ ∞]}∂∂σ}σ=0[σy(z,σ)] ∂ ∑= z σ (σ + n)z n .∂σn=0σ=0The derivative with respect to σ is given by[ ∞]∂ ∑∞∑∑ ∞z σ (σ + n)z n = z σ ln z (σ + n)z n + z σ z n ,∂σn=0n=0n=0which on setting σ = 0 gives the second solution∞∑ ∞∑y 2 (z) =lnz nz n + z nn=0z=(1 − z) ln z + 12 1 − z(z= ln z + 1 )(1 − z) 2 z − 1 .This second solution is the same as that obtained by the Wronskian method in the previoussubsection except <strong>for</strong> the addition of some of the first solution. ◭n=0∞∑16.4.3 Series <strong>for</strong>m of the second solutionUsing any of the methods discussed above, we can find the general <strong>for</strong>m of thesecond solution to the ODE. This <strong>for</strong>m is most easily found, however, using the547


SERIES SOLUTIONS OF ORDINARY DIFFERENTIAL EQUATIONSderivative method. Let us first consider the case where the two solutions of theindicial equation are equal. In this case a second solution is given by (16.28),which may be written as[ ] ∂y(z,σ)y 2 (z) =∂σσ=σ 1∑∞ ∑∞ [ ]=(lnz)z σ1a n (σ 1 )z n dan (σ)+ z σ1dσn=0n=1σ=σ 1z n∑∞= y 1 (z)lnz + z σ1 b n z n , (16.31)n=1where b n =[da n (σ)/dσ] σ=σ1 . One could equally obtain the coefficients b n by directsubstitution of the <strong>for</strong>m (16.31) into the original ODE.In the case where the roots of the indicial equation differ by an integer (notequal to zero), then from (16.30) a second solution is given by{ }∂y 2 (z) =∂σ [(σ − σ 2)y(z,σ)]σ=σ 2[]∑∞=lnz (σ − σ 2 )z σ a n (σ)z n + z σ2n=0σ=σ 2∞ ∑n=0[ ]ddσ (σ − σ 2)a n (σ) z n .σ=σ 2But, as we mentioned in the previous section, [(σ − σ 2 )y(z,σ)] at σ = σ 2 is just amultiple of the first solution y(z,σ 1 ). There<strong>for</strong>e the second solution is of the <strong>for</strong>my 2 (z) =cy 1 (z)lnz + z σ2∞ ∑n=0b n z n , (16.32)where c is a constant. In some cases, however, c might be zero, <strong>and</strong> so the secondsolution would not contain the term in ln z <strong>and</strong> could be written simply as aFrobenius series. Clearly this corresponds to the case in which the substitutionof a Frobenius series into the original ODE yields two solutions automatically.In either case, the coefficients b n may also be found by direct substitution of the<strong>for</strong>m (16.32) into the original ODE.16.5 Polynomial solutionsWe have seen that the evaluation of successive terms of a series solution to adifferential equation is carried out by means of a recurrence relation. The <strong>for</strong>mof the relation <strong>for</strong> a n depends upon n, the previous values of a r (r


16.5 POLYNOMIAL SOLUTIONSis a positive integer or zero, then we are left with a finite polynomial of degreeN ′ = N + σ as a solution of the ODE:N∑y(z) = a n z n+σ . (16.33)n=0In many applications in theoretical physics (particularly in quantum mechanics)the termination of a potentially infinite series after a finite number of termsis of crucial importance in establishing physically acceptable descriptions <strong>and</strong>properties of systems. The condition under which such a termination occurs isthere<strong>for</strong>e of considerable importance.◮Find power series solutions about z =0ofy ′′ − 2zy ′ + λy =0. (16.34)For what values of λ does the equation possess a polynomial solution? Find such a solution<strong>for</strong> λ =4.Clearly z = 0 is an ordinary point of (16.34) <strong>and</strong> so we look <strong>for</strong> solutions of the <strong>for</strong>my = ∑ ∞n=0 a nz n . Substituting this into the ODE <strong>and</strong> multiplying through by z 2 we find∞∑[n(n − 1) − 2z 2 n + λz 2 ]a n z n =0.n=0By dem<strong>and</strong>ing that the coefficients of each power of z vanish separately we derive therecurrence relationn(n − 1)a n − 2(n − 2)a n−2 + λa n−2 =0,which may be rearranged to give2(n − 2) − λa n = a n−2 <strong>for</strong> n ≥ 2. (16.35)n(n − 1)The odd <strong>and</strong> even coefficients are there<strong>for</strong>e independent of one another, <strong>and</strong> two solutionsto (16.34) may be derived. We either set a 1 =0<strong>and</strong>a 0 =1toobtainy 1 (z) =1− λ z2 z4z6− λ(4 − λ) − λ(4 − λ)(8 − λ) − ··· (16.36)2! 4! 6!or set a 0 =0<strong>and</strong>a 1 =1toobtainy 2 (z) =z +(2− λ) z3z5z7+(2− λ)(6 − λ) +(2− λ)(6 − λ)(10 − λ)3! 5! 7! + ··· .Now, from the recurrence relation (16.35) (or in this case from the expressions <strong>for</strong> y 1<strong>and</strong> y 2 themselves) we see that <strong>for</strong> the ODE to possess a polynomial solution we requireλ =2(n − 2) <strong>for</strong> n ≥ 2 or, more simply, λ =2n <strong>for</strong> n ≥ 0, i.e. λ must be an even positiveinteger. If λ = 4 then from (16.36) the ODE has the polynomial solutiony 1 (z) =1− 4z22! =1− 2z2 . ◭A simpler method of obtaining finite polynomial solutions is to assume asolution of the <strong>for</strong>m (16.33), where a N ≠ 0. Instead of starting with the lowestpower of z, as we have done up to now, this time we start by considering the549


SERIES SOLUTIONS OF ORDINARY DIFFERENTIAL EQUATIONScoefficient of the highest power z N ; such a power now exists because of ourassumed <strong>for</strong>m of solution.◮By assuming a polynomial solution find the values of λ in (16.34) <strong>for</strong> which such a solutionexists.We assume a polynomial solution to (16.34) of the <strong>for</strong>m y = ∑ Nn=0 a nz n . Substituting this<strong>for</strong>m into (16.34) we findN∑ [n(n − 1)an z n−2 − 2zna n z n−1 + λa n z n] =0.n=0Now, instead of starting with the lowest power of z, we start with the highest. Thus,dem<strong>and</strong>ing that the coefficient of z N vanishes, we require −2N + λ =0,i.e.λ =2N, aswefound in the previous example. By dem<strong>and</strong>ing that the coefficient of a general power of zis zero, the same recurrence relation as above may be derived <strong>and</strong> the solutions found. ◭16.6 Exercises16.1 Find two power series solutions about z = 0 of the differential equation(1 − z 2 )y ′′ − 3zy ′ + λy =0.Deduce that the value of λ <strong>for</strong> which the corresponding power series becomes anNth-degree polynomial U N (z) isN(N + 2). Construct U 2 (z) <strong>and</strong>U 3 (z).16.2 Find solutions, as power series in z, of the equation4zy ′′ +2(1− z)y ′ − y =0.Identify one of the solutions <strong>and</strong> verify it by direct substitution.16.3 Find power series solutions in z of the differential equationzy ′′ − 2y ′ +9z 5 y =0.Identify closed <strong>for</strong>ms <strong>for</strong> the two series, calculate their Wronskian, <strong>and</strong> verifythat they are linearly independent. Compare the Wronskian with that calculatedfrom the differential equation.16.4 Change the independent variable in the equationd 2 f df+2(z − a) +4f =0dz2 dz (∗)from z to x = z − α, <strong>and</strong> find two independent series solutions, exp<strong>and</strong>ed aboutx = 0, of the resulting equation. Deduce that the general solution of (∗) is∞∑ (−4) m m!f(z,α) =A(z − α)e −(z−α)2 + B(z − α) 2m ,(2m)!m=0with A <strong>and</strong> B arbitrary constants.16.5 Investigate solutions of Legendre’s equation at one of its singular points asfollows.(a) Verify that z = 1 is a regular singular point of Legendre’s equation <strong>and</strong> thatthe indicial equation <strong>for</strong> a series solution in powers of (z − 1) has roots 0<strong>and</strong> 3.(b) Obtain the corresponding recurrence relation <strong>and</strong> show that σ = 0 does notgive a valid series solution.550


16.6 EXERCISES(c) Determine the radius of convergence R of the σ = 3 series <strong>and</strong> relate it tothe positions of the singularities of Legendre’s equation.16.6 Verify that z = 0 is a regular singular point of the equationz 2 y ′′ − 3 2 zy′ +(1+z)y =0,<strong>and</strong> that the indicial equation has roots 2 <strong>and</strong> 1/2. Show that the general solutionis given byy(z) =6a 0 z 2+ b 0(∞∑n=0(−1) n (n +1)2 2n z n(2n +3)!z 1/2 +2z 3/2 − z1/24∞∑n=2)(−1) n 2 2n z n.n(n − 1)(2n − 3)!16.7 Use the derivative method to obtain, as a second solution of Bessel’s equation<strong>for</strong> the case when ν = 0, the following expression:(∞∑n∑)(−1) n 1 ( z) 2nJ 0 (z)lnz −,(n!) 2 r 2n=1r=1given that the first solution is J 0 (z), as specified by (18.79).16.8 Consider a series solution of the equationzy ′′ − 2y ′ + yz =0 (∗)about its regular singular point.(a) Show that its indicial equation has roots that differ by an integer but thatthe two roots nevertheless generate linearly independent solutionsy 1 (z) =3a 0∞∑∞∑n=1(−1) n+1 2nz 2n+1,(2n +1)!(−1) n+1 (2n − 1)z 2ny 2 (z) =a 0.(2n)!n=0(b) Show that y 1 (z) isequalto3a 0 (sin z − z cos z) by exp<strong>and</strong>ing the sinusoidalfunctions. Then, using the Wronskian method, find an expression <strong>for</strong> y 2 (z) interms of sinusoids. You will need to write z 2 as (z/ sin z)(z sin z) <strong>and</strong> integrateby parts to evaluate the integral involved.(c) Confirm that the two solutions are linearly independent by showing thattheir Wronskian is equal to −z 2 , as would be expected from the <strong>for</strong>m of (∗).16.9 Find series solutions of the equation y ′′ − 2zy ′ − 2y = 0. Identify one of the seriesas y 1 (z) =expz 2 <strong>and</strong> verify this by direct substitution. By setting y 2 (z) =u(z)y 1 (z)<strong>and</strong> solving the resulting equation <strong>for</strong> u(z), find an explicit <strong>for</strong>m <strong>for</strong> y 2 (z) <strong>and</strong>deduce that∫ x∑ ∞n!e −v2 dv = e −x202(2n +1)! (2x)2n+1 .n=016.10 Solve the equationz(1 − z) d2 y dy+(1− z) + λy =0dz2 dzas follows.(a) Identify <strong>and</strong> classify its singular points <strong>and</strong> determine their indices.551


SERIES SOLUTIONS OF ORDINARY DIFFERENTIAL EQUATIONS(b) Find one series solution in powers of z. Give a <strong>for</strong>mal expression <strong>for</strong> asecond linearly independent solution.(c) Deduce the values of λ <strong>for</strong> which there is a polynomial solution P N (z) ofdegree N. Evaluate the first four polynomials, normalised in such a way thatP N (0) = 1.16.11 Find the general power series solution about z =0oftheequationz d2 y dy+(2z − 3)dz2 dz + 4 z y =0.16.12 Find the radius of convergence of a series solution about the origin <strong>for</strong> theequation (z 2 + az + b)y ′′ +2y = 0 in the following cases:(a) a =5,b=6; (b)a =5,b=7.Show that if a <strong>and</strong> b are real <strong>and</strong> 4b >a 2 , then the radius of convergence isalways given by b 1/2 .16.13 For the equation y ′′ + z −3 y = 0, show that the origin becomes a regular singularpoint if the independent variable is changed from z to x =1/z. Hence find aseries solution of the <strong>for</strong>m y 1 (z) = ∑ ∞0 a nz −n .Bysettingy 2 (z) =u(z)y 1 (z) <strong>and</strong>exp<strong>and</strong>ing the resulting expression <strong>for</strong> du/dz in powers of z −1 , show that y 2 (z)has the asymptotic <strong>for</strong>m[( )] ln zy 2 (z) =c z +lnz − 1 +O ,2zwhere c is an arbitrary constant.16.14 Prove that the Laguerre equation,z d2 y dy+(1− z) + λy =0,dz2 dzhas polynomial solutions L N (z) ifλ is a non-negative integer N, <strong>and</strong> determinethe recurrence relationship <strong>for</strong> the polynomial coefficients. Hence show that anexpression <strong>for</strong> L N (z), normalised in such a way that L N (0) = N!, isN∑ (−1) n (N!) 2L N (z) =(N − n)!(n!) 2 zn .n=0Evaluate L 3 (z) explicitly.16.15 The origin is an ordinary point of the Chebyshev equation,(1 − z 2 )y ′′ − zy ′ + m 2 y =0,which there<strong>for</strong>e has series solutions of the <strong>for</strong>m z ∑ σ ∞0 a nz n <strong>for</strong> σ =0<strong>and</strong>σ =1.(a) Find the recurrence relationships <strong>for</strong> the a n in the two cases <strong>and</strong> show thatthere exist polynomial solutions T m (z):(i) <strong>for</strong> σ =0,whenm is an even integer, the polynomial having 1 (m +2)2terms;(ii) <strong>for</strong> σ =1,whenm is an odd integer, the polynomial having 1 (m +1)2terms.(b) T m (z) is normalised so as to have T m (1) = 1. Find explicit <strong>for</strong>ms <strong>for</strong> T m (z)<strong>for</strong> m =0, 1, 2, 3.552


16.7 HINTS AND ANSWERS(c)Show that the corresponding non-terminating series solutions S m (z) have astheir first few termsS 0 (z) =a 0(z + 1 3! z3 + 9 )5! z5 + ··· ,S 1 (z) =a 0(1 − 1 2! z2 − 3 )4! z4 − ··· ,(S 2 (z) =a 0 z − 3 3! z3 − 15 )5! z5 − ··· ,S 3 (z) =a 0(1 − 9 2! z2 + 454! z4 + ···16.16 Obtain the recurrence relations <strong>for</strong> the solution of Legendre’s equation (18.1) ininverse powers of z, i.e.sety(z) = ∑ a n z σ−n ,witha 0 ≠ 0. Deduce that, if l is aninteger, then the series with σ = l will terminate <strong>and</strong> hence converge <strong>for</strong> all z,whilst the series with σ = −(l + 1) does not terminate <strong>and</strong> hence converges only<strong>for</strong> |z| > 1.).16.7 Hints <strong>and</strong> answers16.1 Note that z = 0 is an ordinary point of the equation.For σ =0,a n+2 /a n =[n(n +2)− λ]/[(n +1)(n + 2)] <strong>and</strong>, correspondingly, <strong>for</strong>σ =1,U 2 (z) =a 0 (1 − 4z 2 )<strong>and</strong>U 3 (z) =a 0 (z − 2z 3 ).16.3 σ =0<strong>and</strong>3;a 6m /a 0 =(−1) m /(2m)! <strong>and</strong> a 6m /a 0 =(−1) m /(2m + 1)!, respectively.y 1 (z) =a 0 cos z 3 <strong>and</strong> y 2 (z) =a 0 sin z 3 .TheWronskianis±3a 2 0 z2 ≠0.16.5 (b) a n+1 /a n =[l(l +1)− n(n +1)]/[2(n +1) 2 ].(c) R = 2, equal to the distance between z = 1 <strong>and</strong> the closest singularity atz = −1.(−1) n z 2n16.7 A typical term in the series <strong>for</strong> y(σ, z) is[(σ +2)(σ +4)···(σ +2n)] . 216.9 The origin is an ordinary point. Determine the constant of integration by examiningthe behaviour of the related functions <strong>for</strong> small x.y 2 (z) =(expz 2 ) ∫ z0 exp(−x2 ) dx.16.11 Repeated roots σ =2.∞∑y(z) =az 2 − 4az 3 +6bz 3 (n +1)(−2z) n+2 { a}+n! 4 + b [ln z + g(n)] ,n=2whereg(n) = 1n +1 − 1 n − 1n − 1 − ···− 1 2 − 2.16.13 The trans<strong>for</strong>med equation is xy ′′ +2y ′ + y =0;a n =(−1) n (n +1) −1 (n!) −2 a 0 ;du/dz = A[ y 1 (z)] −2 .16.15 (a) (i) a n+2 =[a n (n 2 − m 2 )]/[(n +2)(n + 1)],(ii) a n+2 = {a n [(n +1) 2 − m 2 ]}/[(n +3)(n + 2)]; (b) 1, z, 2z 2 − 1, 4z 3 − 3z.553


17Eigenfunction methods <strong>for</strong>differential equationsIn the previous three chapters we dealt with the solution of differential equationsof order n by two methods. In one method, we found n independent solutionsof the equation <strong>and</strong> then combined them, weighted with coefficients determinedby the boundary conditions; in the other we found solutions in terms of serieswhose coefficients were related by (in general) an n-term recurrence relation <strong>and</strong>thence fixed by the boundary conditions. For both approaches the linearity of theequation was an important or essential factor in the utility of the method, <strong>and</strong>in this chapter our aim will be to exploit the superposition properties of lineardifferential equations even further.We will be concerned with the solution of equations of the inhomogeneous<strong>for</strong>mLy(x) =f(x), (17.1)where f(x) is a prescribed or general function <strong>and</strong> the boundary conditions tobe satisfied by the solution y = y(x), <strong>for</strong> example at the limits x = a <strong>and</strong> x = b,are given. The expression Ly(x) st<strong>and</strong>s <strong>for</strong> a linear differential operator L actingupon the function y(x).In general, unless f(x) is both known <strong>and</strong> simple, it will not be possible tofind particular integrals of (17.1), even if complementary functions can be foundthat satisfy Ly = 0. The idea is there<strong>for</strong>e to exploit the linearity of L by buildingup the required solution y(x) asasuperposition, generally containing an infinitenumber of terms, of some set of functions {y i (x)} that each individually satisfythe boundary conditions. Clearly this brings in a quite considerable complicationbut since, within reason, we may select the set of functions to suit ourselves, wecan obtain sizeable compensation <strong>for</strong> this complication. Indeed, if the set chosenis one containing functions that, when acted upon by L, produce particularlysimple results then we can ‘show a profit’ on the operation. In particular, if the554


EIGENFUNCTION METHODS FOR DIFFERENTIAL EQUATIONSset consists of those functions y i <strong>for</strong> whichLy i (x) =λ i y i (x), (17.2)where λ i is a constant (<strong>and</strong> which satisfy the boundary conditions), then a distinctadvantage may be obtained from the manoeuvre because all the differentiationwill have disappeared from (17.1).Equation (17.2) is clearly reminiscent of the equation satisfied by the eigenvectorsx i of a linear operator A , namelyA x i = λ i x i , (17.3)where λ i is a constant <strong>and</strong> is called the eigenvalue associated with x i . By analogy,in the context of differential equations a function y i (x) satisfying (17.2) is calledan eigenfunction of the operator L (under the imposed boundary conditions) <strong>and</strong>λ i is then called the eigenvalue associated with the eigenfunction y i (x). Clearly,the eigenfunctions y i (x) ofL are only determined up to an arbitrary scale factorby (17.2).Probably the most familiar equation of the <strong>for</strong>m (17.2) is that which describesa simple harmonic oscillator, i.e.Ly ≡− d2 ydt 2 = ω2 y, where L ≡−d 2 /dt 2 . (17.4)Imposing the boundary condition that the solution is periodic with period T ,the eigenfunctions in this case are given by y n (t) =A n e iωnt ,whereω n =2πn/T ,n =0, ±1, ±2,... <strong>and</strong> the A n are constants. The eigenvalues are ωn 2 = n 2 ω1 2 =n 2 (2π/T) 2 . (Sometimes ω n is referred to as the eigenvalue of this equation, butwe will avoid such confusing terminology here.)We may discuss a somewhat wider class of differential equations by consideringa slightly more general <strong>for</strong>m of (17.2), namelyLy i (x) =λ i ρ(x)y i (x), (17.5)where ρ(x) isaweight function. In many applications ρ(x) is unity <strong>for</strong> all x, inwhich case (17.2) is recovered; in general, though, it is a function determined bythe choice of coordinate system used in describing a particular physical situation.The only requirement on ρ(x) is that it is real <strong>and</strong> does not change sign in therange a ≤ x ≤ b, so that it can, without loss of generality, be taken to be nonnegativethroughout; of course, ρ(x) must be the same function <strong>for</strong> all values ofλ i . A function y i (x) that satisfies (17.5) is called an eigenfunction of the operatorL with respect to the weight function ρ(x).This chapter will not cover methods used to determine the eigenfunctions of(17.2) or (17.5), since we have discussed those in previous chapters, but, rather, willuse the properties of the eigenfunctions to solve inhomogeneous equations of the<strong>for</strong>m (17.1). We shall see later that the sets of eigenfunctions y i (x) of a particular555


EIGENFUNCTION METHODS FOR DIFFERENTIAL EQUATIONSclass of operators called Hermitian operators (the operator in the simple harmonicoscillator equation is an example) have particularly useful properties <strong>and</strong> thesewill be studied in detail. It turns out that many of the interesting differentialoperators met within the physical sciences are Hermitian. Be<strong>for</strong>e continuing ourdiscussion of the eigenfunctions of Hermitian operators, however, we will considersome properties of general sets of functions.17.1 Sets of functionsIn chapter 8 we discussed the definition of a vector space but concentrated onspaces of finite dimensionality. We consider now the infinite-dimensional spaceof all reasonably well-behaved functions f(x), g(x), h(x), ... on the intervala ≤ x ≤ b. That these functions <strong>for</strong>m a linear vector space is shown by notingthe following properties. The set is closed under(i) addition, which is commutative <strong>and</strong> associative, i.e.f(x)+g(x) =g(x)+f(x),[f(x)+g(x)] + h(x) =f(x)+[g(x)+h(x)] ,(ii) multiplication by a scalar, which is distributive <strong>and</strong> associative, i.e.Furthermore, in such a spaceλ [f(x)+g(x)] = λf(x)+λg(x),λ [µf(x)] =(λµ)f(x),(λ + µ)f(x) =λf(x)+µf(x).(iii) there exists a ‘null vector’ 0 such that f(x)+0=f(x),(iv) multiplication by unity leaves any function unchanged, i.e. 1 × f(x) =f(x),(v) each function has an associated negative function −f(x) that is such thatf(x)+[−f(x)] = 0.By analogy with finite-dimensional vector spaces we now introduce a setof linearly independent basis functions y n (x), n = 0, 1,...,∞, such that any‘reasonable’ function in the interval a ≤ x ≤ b (i.e. it obeys the Dirichlet conditionsdiscussed in chapter 12) can be expressed as the linear sum of these functions:∞∑f(x) = c n y n (x).n=0Clearly if a different set of linearly independent basis functions u n (x) is chosenthen the function can be expressed in terms of the new basis,∞∑f(x) = d n u n (x),n=0556


17.1 SETS OF FUNCTIONSwhere the d n are a different set of coefficients. In each case, provided the basisfunctions are linearly independent, the coefficients are unique.We may also define an inner product on our function space by〈f|g〉 =∫ baf ∗ (x)g(x)ρ(x) dx, (17.6)where ρ(x) is the weight function, which we require to be real <strong>and</strong> non-negativein the interval a ≤ x ≤ b. As mentioned above, ρ(x) is often unity <strong>for</strong> all x. Twofunctions are said to be orthogonal (with respect to the weight function ρ(x)) onthe interval [a, b] if〈f|g〉 =∫ baf ∗ (x)g(x)ρ(x) dx =0, (17.7)<strong>and</strong> the norm of a function is defined as[∫ b1/2 [∫ b1/2‖f‖ = 〈f|f〉 1/2 = f ∗ (x)f(x)ρ(x) dx]= |f(x)| 2 ρ(x) dx]. (17.8)aIt is also common practice to define a normalised function by ˆf = f/‖f‖, whichhas unit norm.An infinite-dimensional vector space of functions, <strong>for</strong> which an inner productis defined, is called a Hilbert space. Using the concept of the inner product, wecan choose a basis of linearly independent functions ˆφ n (x), n =0, 1, 2,... that areorthonormal, i.e. such that〈 ˆφ i | ˆφ j 〉 =∫ baaˆφ ∗ i (x) ˆφ j (x)ρ(x) dx = δ ij . (17.9)If y n (x), n =0, 1, 2,..., are a linearly independent, but not orthonormal, basis <strong>for</strong>the Hilbert space then an orthonormal set of basis functions ˆφ n may be produced(in a similar manner to that used in the construction of a set of orthogonaleigenvectors of an Hermitian matrix; see chapter 8) by the following procedure:φ 0 = y 0 ,φ 1 = y 1 − ˆφ 0 〈 ˆφ 0 |y 1 〉,φ 2 = y 2 − ˆφ 1 〈 ˆφ 1 |y 2 〉− ˆφ 0 〈 ˆφ 0 |y 2 〉,.φ n = y n − ˆφ n−1 〈 ˆφ n−1 |y n 〉−···− ˆφ 0 〈 ˆφ 0 |y n 〉,.It is straight<strong>for</strong>ward to check that each φ n is orthogonal to all its predecessorsφ i , i =0, 1, 2,...,n− 1. This method is called Gram–Schmidt orthogonalisation.Clearly the functions φ n <strong>for</strong>m an orthogonal set, but in general they do not haveunit norms.557


EIGENFUNCTION METHODS FOR DIFFERENTIAL EQUATIONS◮Starting from the linearly independent functions y n (x) =x n , n =0, 1,..., construct threeorthonormal functions over the range −1


17.2 ADJOINT, SELF-ADJOINT AND HERMITIAN OPERATORS17.1.1 Some useful inequalitiesSince <strong>for</strong> a Hilbert space 〈f|f〉 ≥0, the inequalities discussed in subsection 8.1.3hold. The proofs are not repeated here, but the relationships are listed <strong>for</strong>completeness.(i) The Schwarz inequality states that|〈f|g〉| ≤ 〈f|f〉 1/2 〈g|g〉 1/2 , (17.12)where the equality holds when f(x) is a scalar multiple of g(x), i.e. whenthey are linearly dependent.(ii) The triangle inequality states that‖f + g‖ ≤‖f‖ + ‖g‖, (17.13)where again equality holds when f(x) is a scalar multiple of g(x).(iii) Bessel’s inequality requires the introduction of an orthonormal basis ˆφ n (x)so that any function f(x) can be written as∞∑f(x) = c n ˆφn (x),n=0where c n = 〈 ˆφ n |f〉. Bessel’s inequality then states that〈f|f〉 ≥ ∑ n|c n | 2 . (17.14)The equality holds if the summation is over all the basis functions. If somevalues of n are omitted from the sum then the inequality results (unless,of course, the c n happen to be zero <strong>for</strong> all values of n omitted, in whichcase the equality remains).17.2 Adjoint, self-adjoint <strong>and</strong> Hermitian operatorsHaving discussed general sets of functions, we now return to the discussion ofeigenfunctions of linear operators. We begin by introducing the adjoint of anoperator L, denoted by L † , which is defined by∫ baf ∗ (x) [Lg(x)] dx =∫ ba[L † f(x)] ∗ g(x) dx + boundary terms,(17.15)where the boundary terms are evaluated at the end-points of the interval [a, b].Thus, <strong>for</strong> any given linear differential operator L, the adjoint operator L † can befound by repeated integration by parts.An operator is said to be self-adjoint if L † = L. If, in addition, certain boundaryconditions are met by the functions f <strong>and</strong> g on which a self-adjoint operator acts,559


EIGENFUNCTION METHODS FOR DIFFERENTIAL EQUATIONSor by the operator itself, such that the boundary terms in (17.15) vanish, then theoperator is said to be Hermitian over the interval a ≤ x ≤ b. Thus, in this case,∫ baf ∗ (x) [Lg(x)] dx =∫ ba[Lf(x)] ∗ g(x) dx. (17.16)A little careful study will reveal the similarity between the definition of anHermitian operator <strong>and</strong> the definition of an Hermitian matrix given in chapter 8.◮Show that the linear operator L = d 2 /dt 2 is self-adjoint, <strong>and</strong> determine the requiredboundary conditions <strong>for</strong> the operator to be Hermitian over the interval t 0 to t 0 + T .Substituting into the LHS of the definition of the adjoint operator (17.15) <strong>and</strong> integratingby parts gives∫ t0 +T[f ∗ d2 gt 0dt dt = f ∗ dg ] t0 +T ∫ t0 +Tdf ∗ dg−2 dtt 0 t 0dt dt dt.Integrating the second term on the RHS by parts once more yields∫ t0 +T[f ∗ d2 gt 0dt dt = f ∗ dg ] t0 +T ] t0+[− df∗ +T ∫ t0 +T2 dtt 0dt g + g d2 f ∗t 0 t 0dt dt, 2which, by comparison with (17.15), proves that L is a self-adjoint operator. Moreover,from (17.16), we see that L is an Hermitian operator over the required interval provided[f ∗ dg ] t0 +T [ ] df∗ t0 +Tdt dt g . ◭t 0=We showed in chapter 8 that the eigenvalues of Hermitian matrices are real <strong>and</strong>that their eigenvectors can be chosen to be orthogonal. Similarly, the eigenvaluesof Hermitian operators are real <strong>and</strong> their eigenfunctions can be chosen to beorthogonal (we will prove these properties in the following section). Hermitianoperators (or matrices) are often used in the <strong>for</strong>mulation of quantum mechanics.The eigenvalues then give the possible measured values of an observable quantitysuch as energy or angular momentum, <strong>and</strong> the physical requirement that suchquantities must be real is ensured by the reality of these eigenvalues. Furthermore,the infinite set of eigenfunctions of an Hermitian operator <strong>for</strong>m a complete basisset over the relevant interval, so that it is possible to exp<strong>and</strong> any function y(x)obeying the appropriate conditions in an eigenfunction series over this interval:∞∑y(x) = c n y n (x), (17.17)n=0where the choice of suitable values <strong>for</strong> the c n will make the sum arbitrarily closeto y(x). § These useful properties provide the motivation <strong>for</strong> a detailed study ofHermitian operators.t 0§ The proof of the completeness of the eigenfunctions of an Hermitian operator is beyond the scopeof this book. The reader should refer, <strong>for</strong> example, to R. Courant <strong>and</strong> D. Hilbert, <strong>Methods</strong> of<strong>Mathematical</strong> <strong>Physics</strong> (New York: Interscience, 1953).560


17.3 PROPERTIES OF HERMITIAN OPERATORS17.3 Properties of Hermitian operatorsWe now provide proofs of some of the useful properties of Hermitian operators.Again much of the analysis is similar to that <strong>for</strong> Hermitian matrices in chapter 8,although the present section st<strong>and</strong>s alone. (Here, <strong>and</strong> throughout the remainderof this chapter, we will write out inner products in full. We note, however, thatthe inner product notation often provides a neat <strong>for</strong>m in which to express results.)17.3.1 Reality of the eigenvaluesConsider an Hermitian operator <strong>for</strong> which (17.5) is satisfied by at least twoeigenfunctions y i (x) <strong>and</strong>y j (x), which have corresponding eigenvalues λ i <strong>and</strong> λ j ,so thatLy i = λ i ρ(x)y i , (17.18)Ly j = λ j ρ(x)y j , (17.19)where we have allowed <strong>for</strong> the presence of a weight function ρ(x). Multiplying(17.18) by yj ∗ <strong>and</strong> (17.19) by y∗ i <strong>and</strong> then integrating gives∫ b∫ byj ∗ Ly i dx = λ i yj ∗ y i ρdx, (17.20)a∫ baa∫ byi ∗ Ly j dx = λ j yi ∗ y j ρdx. (17.21)Remembering that we have required ρ(x) to be real, the complex conjugate of(17.20) becomes∫ bay j (Ly i ) ∗ dx = λ ∗ ia∫ bay ∗ i y j ρdx, (17.22)<strong>and</strong> using the definition of an Hermitian operator (17.16) it follows that the LHSof (17.22) is equal to the LHS of (17.21). Thus(λ ∗ i − λ j )∫ bay ∗ i y j ρdx=0. (17.23)If i = j then λ i = λ ∗ i (since ∫ ba y∗ i y iρdx ≠ 0), which is a statement that theeigenvalue λ i is real.17.3.2 Orthogonality <strong>and</strong> normalisation of the eigenfunctionsFrom (17.23), it is immediately apparent that two eigenfunctions y i <strong>and</strong> y j thatcorrespond to different eigenvalues, i.e. such that λ i ≠ λ j , satisfy∫ bay ∗ i y j ρdx=0, (17.24)561


EIGENFUNCTION METHODS FOR DIFFERENTIAL EQUATIONSwhich is a statement of the orthogonality of y i <strong>and</strong> y j .If one (or more) of the eigenvalues is degenerate, however, we have differenteigenfunctions corresponding to the same eigenvalue, <strong>and</strong> the proof of orthogonalityis not so straight<strong>for</strong>ward. Nevertheless, an orthogonal set of eigenfunctionsmay be constructed using the Gram–Schmidt orthogonalisation method mentionedearlier in this chapter <strong>and</strong> used in chapter 8 to construct a set of orthogonaleigenvectors of an Hermitian matrix. We repeat the analysis here <strong>for</strong> completeness.Suppose, <strong>for</strong> the sake of our proof, that λ 0 is k-fold degenerate, i.e.Ly i = λ 0 ρy i <strong>for</strong> i =0, 1,...,k− 1, (17.25)but that λ 0 is different from any of λ k , λ k+1 , etc. Then any linear combination ofthese y i is also an eigenfunction with eigenvalue λ 0 since∑k−1∑k−1∑k−1Lz ≡ L c i y i = c i Ly i = c i λ 0 ρy i = λ 0 ρz. (17.26)i=0i=0If the y i defined in (17.25) are not already mutually orthogonal then considerthe new eigenfunctions z i constructed by the following procedure, in which eachof the new functions z i is to be normalised, to give ẑ i , be<strong>for</strong>e proceeding to theconstruction of the next one (the normalisation can be carried out by dividingthe eigenfunction z i by ( ∫ ba z∗ i z iρdx) 1/2 ):z 0 = y 0 ,∫ b)z 1 = y 1 −(ẑ 0 ẑ0y ∗ 1 ρdx ,a∫ b) ∫ b)z 2 = y 2 −(ẑ 1 ẑ1y ∗ 2 ρdx −(ẑ 0 ẑ0y ∗ 2 ρdx ,aa.∫ b)∫ b)z k−1 = y k−1 −(ẑ k−2 ẑk−2y ∗ k−1 ρdx − ···−(ẑ 0 ẑ0y ∗ k−1 ρdx .aaEach of the integrals is just a number <strong>and</strong> thus each new function z i is, as can beshown from (17.26), an eigenvector of L with eigenvalue λ 0 . It is straight<strong>for</strong>wardto check that each z i is orthogonal to all its predecessors. Thus, by this explicitconstruction we have shown that an orthogonal set of eigenfunctions of anHermitian operator L can be obtained. Clearly the orthogonal set obtained, z i ,isnot unique.In general, since L is linear, the normalisation of its eigenfunctions y i (x) isarbitrary. It is often convenient, however, to work in terms of the normalisedeigenfunctions ŷ i (x), so that ∫ ba ŷ∗ i ŷiρdx= 1. These there<strong>for</strong>e <strong>for</strong>m an orthonormal562i=0


17.3 PROPERTIES OF HERMITIAN OPERATORSset <strong>and</strong> we can write∫ bwhich is valid <strong>for</strong> all pairs of values i, j.aŷ ∗ i ŷjρdx= δ ij , (17.27)17.3.3 Completeness of the eigenfunctionsAs noted earlier, the eigenfunctions of an Hermitian operator may be shownto <strong>for</strong>m a complete basis set over the relevant interval. One may thus exp<strong>and</strong>any (reasonable) function y(x) obeying appropriate boundary conditions in aneigenfunction series over the interval, as in (17.17). Working in terms of thenormalised eigenfunctions ŷ n (x), we may thus writef(x) = ∑ n∫ bŷ n (x)∫ baŷ ∗ n(z)f(z)ρ(z) dz= f(z)ρ(z) ∑ ŷ n (x)ŷn(z) ∗ dz.anSince this is true <strong>for</strong> any f(x), we must have thatρ(z) ∑ nŷ n (x)ŷ ∗ n(z) =δ(x − z). (17.28)This is called the completeness or closure property of the eigenfunctions. It definesa complete set. If the spectrum of eigenvalues of L is anywhere continuous thenthe eigenfunction y n (x) must be treated as y(n, x) <strong>and</strong> an integration carried outover n.We also note that the RHS of (17.28) is a δ-function <strong>and</strong> so is only non-zerowhen z = x; thus ρ(z) on the LHS can be replaced by ρ(x) if required, i.e.ρ(z) ∑ nŷ n (x)ŷ ∗ n(z) =ρ(x) ∑ nŷ n (x)ŷ ∗ n(z). (17.29)17.3.4 Construction of real eigenfunctionsRecall that the eigenfunction y i satisfiesLy i = λ i ρy i (17.30)<strong>and</strong> that the complex conjugate of this givesLyi ∗ = λ ∗ i ρyi ∗ = λ i ρyi ∗ , (17.31)where the last equality follows because the eigenvalues are real, i.e. λ i = λ ∗ i .Thus, y i <strong>and</strong> yi∗ are eigenfunctions corresponding to the same eigenvalue <strong>and</strong>hence, because of the linearity of L, atleastoneofyi ∗ + y i <strong>and</strong> i(yi ∗ − y i ), which563


EIGENFUNCTION METHODS FOR DIFFERENTIAL EQUATIONSare both real, is a non-zero eigenfunction corresponding to that eigenvalue. Itfollows that the eigenfunctions can always be made real by taking suitable linearcombinations, though taking such linear combinations will only be necessary incases where a particular λ is degenerate, i.e. corresponds to more than one linearlyindependent eigenfunction.17.4 Sturm–Liouville equationsOne of the most important applications of our discussion of Hermitian operatorsis to the study of Sturm–Liouville equations, which take the general <strong>for</strong>mp(x) d2 y dy+ r(x) + q(x)y + λρ(x)y =0,dx2 dxwhere r(x) =dp(x)dx(17.32)<strong>and</strong> p, q <strong>and</strong> r are real functions of x. § A variational approach to the Sturm–Liouville equation, which is useful in estimating the eigenvalues λ <strong>for</strong> a given setof boundary conditions on y, is discussed in chapter 22. For now, however, weconcentrate on demonstrating that solutions of the Sturm–Liouville equation thatsatisfy appropriate boundary conditions are the eigenfunctions of an Hermitianoperator.It is clear that (17.32) can be writtenLy = λρ(x)y, where L ≡−[p(x) d2dx 2 + r(x) d ]dx + q(x) . (17.33)Using the condition that r(x) =p ′ (x), it will be seen that the general Sturm–Liouville equation (17.32) can also be rewritten as(py ′ ) ′ + qy + λρy =0, (17.34)where primes denote differentiation with respect to x. Using (17.33) this may alsobe written Ly ≡−(py ′ ) ′ − qy = λρy, which defines a more useful <strong>for</strong>m <strong>for</strong> theSturm–Liouville linear operator, namelyL ≡−[ ddx(p(x) ddx) ]+ q(x) . (17.35)17.4.1 Hermitian nature of the Sturm–Liouville operatorAs we now show, the linear operator of the Sturm–Liouville equation (17.35) isself-adjoint. Moreover, the operator is Hermitian over the range [a, b] provided§ We note that sign conventions vary in this expression <strong>for</strong> the general Sturm–Liouville equation;some authors use −λρ(x)y on the LHS of (17.32).564


17.4 STURM–LIOUVILLE EQUATIONScertain boundary conditions are met, namely that any two eigenfunctions y i <strong>and</strong>y j of (17.33) must satisfy[y∗i py j]′ y ∗ x=a i py j′ ]x=b<strong>for</strong> all i, j. (17.36)Rearranging (17.36), we can write[ ] x=byi ∗ py j′ = 0 x=a (17.37)as an equivalent statement of the required boundary conditions. These boundaryconditions are in fact not too restrictive <strong>and</strong> are met, <strong>for</strong> instance, by the setsy(a) =y(b) =0;y(a) =y ′ (b) =0;p(a) =p(b) = 0 <strong>and</strong> by many other sets. Itis important to note that in order to satisfy (17.36) <strong>and</strong> (17.37) one boundarycondition must be specified at each end of the range.◮Prove that the Sturm–Liouville operator is Hermitian over the range [a, b] <strong>and</strong> under theboundary conditions (17.37).Putting the Sturm–Liouville <strong>for</strong>m Ly = −(py ′ ) ′ − qy into the definition (17.16) of anHermitian operator, the LHS may be written as a sum of two terms, i.e.∫ b∫[ ] b∫ b− y∗i (py j) ′ ′ + yi ∗ qy j dx = − yi ∗ (py j) ′ ′ dx − yi ∗ qy j dx.aThe first term may be integrated by parts to give[ ] b ∫ b− yi ∗ py j′ + (yi ∗ ) ′ py j ′ dx.a aThe boundary-value term in this is zero because of the boundary conditions, <strong>and</strong> sointegrating by parts again yields[ ] b ∫ b(yi ∗ ) ′ py j − ((yi ∗ ) ′ p) ′ y j dx.aAgain, the boundary-value term is zero, leaving us with−∫ ba[y∗i (py ′ j) ′ + y ∗ i qy j]dx = −∫ baaa[ ]yj (p(yi ∗ ) ′ ) ′ + y j qyi∗ dx,which proves that the Sturm–Liouville operator is Hermitian over the prescribed interval. ◭It is also worth noting that, since p(a) =p(b) = 0 is a valid set of boundaryconditions, many Sturm–Liouville equations possess a ‘natural’ interval [a, b] overwhich the corresponding differential operator L is Hermitian irrespective of theboundary conditions satisfied by its eigenfunctions at x = a <strong>and</strong> x = b (the onlyrequirement being that they are regular at these end-points).a17.4.2 Trans<strong>for</strong>ming an equation into Sturm–Liouville <strong>for</strong>mMany of the second-order differential equations encountered in physical problemsare examples of the Sturm–Liouville equation (17.34). Moreover, any second-order565


EIGENFUNCTION METHODS FOR DIFFERENTIAL EQUATIONSEquation p(x) q(x) λ ρ(x)Hypergeometric x c (1 − x) a+b−c+1 0 −ab x c−1 (1 − x) a+b−cLegendre 1 − x 2 0 l(l +1) 1Associated Legendre 1 − x 2 −m 2 /(1 − x 2 ) l(l +1) 1Chebyshev (1 − x 2 ) 1/2 0 ν 2 (1 − x 2 ) −1/2Confluent hypergeometric x c e −x 0 −a x c−1 e −xBessel ∗ x −ν 2 /x α 2 xLaguerre xe −x 0 ν e −xAssociated Laguerre x m+1 e −x 0 ν x m e −xHermite e −x2 0 2ν e −x2Simple harmonic 1 0 ω 2 1Table 17.1 The Sturm–Liouville <strong>for</strong>m (17.34) <strong>for</strong> important ODEs in thephysical sciences <strong>and</strong> engineering. The asterisk denotes that, <strong>for</strong> Bessel’s equation,a change of variable x → x/a is required to give the conventionalnormalisation used here, but is not needed <strong>for</strong> the trans<strong>for</strong>mation into Sturm–Liouville <strong>for</strong>m.differential equation of the <strong>for</strong>mp(x)y ′′ + r(x)y ′ + q(x)y + λρ(x)y = 0 (17.38)can be converted into Sturm–Liouville <strong>for</strong>m by multiplying through by a suitableintegrating factor, which is given by{∫ xr(u) − p ′ }(u)F(x) =expdu . (17.39)p(u)It is easily verified that (17.38) then takes the Sturm–Liouville <strong>for</strong>m,[F(x)p(x)y ′ ] ′ + F(x)q(x)y + λF(x)ρ(x)y =0, (17.40)with a different, but still non-negative, weight function F(x)ρ(x). Table 17.1summarises the Sturm–Liouville <strong>for</strong>m (17.34) <strong>for</strong> several of the equations listedin table 16.1. These <strong>for</strong>ms can be determined using (17.39), as illustrated in thefollowing example.◮Put the following equations into Sturm–Liouville (SL) <strong>for</strong>m:(i) (1 − x 2 )y ′′ − xy ′ + ν 2 y =0 (Chebyshev equation);(ii) xy ′′ +(1− x)y ′ + νy =0 (Laguerre equation);(iii) y ′′ − 2xy ′ +2νy =0 (Hermite equation).(i) From (17.39), the required integrating factor is(∫ x) uF(x) =exp1 − u du =exp [ − 1 ln(1 − 2 2 x2 ) ] =(1− x 2 ) −1/2 .Thus, the Chebyshev equation becomes(1 − x 2 ) 1/2 y ′′ − x(1 − x 2 ) −1/2 y ′ + ν 2 (1 − x 2 ) −1/2 y = [ (1 − x 2 ) 1/2 y ′] ′+ ν 2 (1 − x 2 ) −1/2 y =0,whichisinSL<strong>for</strong>mwithp(x) =(1− x 2 ) 1/2 , q(x) =0,ρ(x) =(1− x 2 ) −1/2 <strong>and</strong> λ = ν 2 .566


17.4 STURM–LIOUVILLE EQUATIONS(ii) From (17.39), the required integrating factor isx)F(x) =exp(∫−1 du =exp(−x).Thus, the Laguerre equation becomesxe −x y ′′ +(1− x)e −x y ′ + νe −x y =(xe −x y ′ ) ′ + νe −x y =0,which is in SL <strong>for</strong>m with p(x) =xe −x , q(x) =0,ρ(x) =e −x <strong>and</strong> λ = ν.(iii) From (17.39), the required integrating factor isx)F(x) =exp(∫−2udu =exp(−x 2 ).Thus, the Hermite equation becomese −x2 y ′′ − 2xe −x2 y ′ +2νe −x2 y =(e −x2 y ′ ) ′ +2νe −x2 y =0,which is in SL <strong>for</strong>m with p(x) =e −x2 , q(x) =0,ρ(x) =e −x2 <strong>and</strong> λ =2ν. ◭From the p(x) entries in table 17.1, we may read off the natural interval overwhich the corresponding Sturm–Liouville operator (17.35) is Hermitian; in eachcase this is given by [a, b], where p(a) =p(b) = 0. Thus, the natural interval<strong>for</strong> the Legendre equation, the associated Legendre equation <strong>and</strong> the Chebyshevequation is [−1, 1]; <strong>for</strong> the Laguerre <strong>and</strong> associated Laguerre equations theinterval is [0, ∞]; <strong>and</strong> <strong>for</strong> the Hermite equation it is [−∞, ∞]. In addition, from(17.37), one sees that <strong>for</strong> the simple harmonic equation one requires only that[a, b] =[x 0 ,x 0 +2π]. We also note that, as required, the weight function in eachcase is finite <strong>and</strong> non-negative over the natural interval. Occasionally, a little morecare is required when determining the conditions <strong>for</strong> a Sturm–Liouville operatorof the <strong>for</strong>m (17.35) to be Hermitian over some natural interval, as is illustratedin the following example.◮Express the hypergeometric equation,x(1 − x)y ′′ +[c − (a + b +1)x ]y ′ − aby =0,in Sturm–Liouville <strong>for</strong>m. Hence determine the natural interval over which the resultingSturm–Liouville operator is Hermitian <strong>and</strong> the corresponding conditions that one must imposeon the parameters a, b <strong>and</strong> c.As usual <strong>for</strong> an equation not already in SL <strong>for</strong>m, we first determine the appropriate567


EIGENFUNCTION METHODS FOR DIFFERENTIAL EQUATIONSintegrating factor. This is given, as in equation (17.39), by[ ∫ xc − (a + b +1)u − 1+2uF(x) =expu(1 − u)[ ∫ x]c − 1 − (a + b − 1)u=expduu(1 − u)[ ∫ x( c − 1=exp1 − u + c − 1u]du− a + b − 1 )]du1 − u=exp[(a + b − c)ln(1− x)+(c − 1) ln x ]= x c−1 (1 − x) a+b−c .When the equation is multiplied through by F(x) ittakesthe<strong>for</strong>m[x c (1 − x) a+b−c+1 y ′ ] ′− abx c−1 (1 − x) a+b−c y =0.Now, <strong>for</strong> the corresponding Sturm–Liouville operator to be Hermitian, the conditionsto be imposed are as follows.(i) The boundary condition (17.37); if c>0<strong>and</strong>a + b − c +1> 0, this is satisfiedautomatically <strong>for</strong> 0 ≤ x ≤ 1, which is thus the natural interval in this case.(ii) The weight function x c−1 (1 − x) a+b−c must be finite <strong>and</strong> not change sign in theinterval 0 ≤ x ≤ 1. This means that both exponents in it must be positive, i.e.c − 1 > 0<strong>and</strong>a + b − c>0.Putting together the conditions on the parameters gives the double inequality a + b>c>1. ◭Finally, we consider Bessel’s equation,x 2 y ′′ + xy ′ +(x 2 − ν 2 )y =0,which may be converted into Sturm–Liouville <strong>for</strong>m, but only in a somewhatunorthodox fashion. It is conventional first to divide the Bessel equation by x<strong>and</strong> then to change variables to ¯x = x/α. In this case, it becomes¯xy ′′ (α¯x)+y ′ (α¯x) − ν2¯x y(α¯x)+α2¯xy(α¯x) =0, (17.41)where a prime now indicates differentiation with respect to ¯x. Dropping the barson the independent variable, we thus have[xy ′ (αx)] ′ − ν2x y(αx)+α2 xy(αx) =0, (17.42)whichisinSL<strong>for</strong>mwithp(x) =x, q(x) =−ν 2 /x, ρ(x) =x <strong>and</strong> λ = α 2 .Itshould be noted, however, that in this case the eigenvalue (actually its squareroot) appears in the argument of the dependent variable.568


17.5 SUPERPOSITION OF EIGENFUNCTIONS: GREEN’S FUNCTIONS17.5 Superposition of eigenfunctions: Green’s functionsWe have already seen that ifLy n (x) =λ n ρ(x)y n (x), (17.43)where L is an Hermitian operator, then the eigenvalues λ n are real <strong>and</strong> theeigenfunctions y n (x) are orthogonal (or can be made so). Let us assume that weknow the eigenfunctions y n (x) ofL that individually satisfy (17.43) <strong>and</strong> someimposed boundary conditions (<strong>for</strong> which L is Hermitian).Now let us suppose we wish to solve the inhomogeneous differential equationLy(x) =f(x), (17.44)subject to the same boundary conditions. Since the eigenfunctions of L <strong>for</strong>m acomplete set, the full solution, y(x), to (17.44) may be written as a superpositionof eigenfunctions, i.e.∞∑y(x) = c n y n (x), (17.45)n=0<strong>for</strong> some choice of the constants c n . Making full use of the linearity of L, we have( ∞)∑∞∑∞∑f(x) =Ly(x) =L c n y n (x) = c n Ly n (x) = c n λ n ρ(x)y n (x).n=0n=0n=0(17.46)Multiplying the first <strong>and</strong> last terms of (17.46) by yj ∗ <strong>and</strong> integrating, we obtain∫ bay ∗ j (z)f(z) dz =∞∑n=0∫ bac n λ n y ∗ j (z)y n (z)ρ(z) dz, (17.47)where we have used z as the integration variable <strong>for</strong> later convenience. Finally,using the orthogonality condition (17.27), we see that the integrals on the RHSare zero unless n = j, <strong>and</strong> so obtainc n = 1 ∫ ba y∗ n(z)f(z) dz∫λ b na y∗ n (z)y n(z)ρ(z) dz . (17.48)Thus, if we can find all the eigenfunctions of a differential operator then (17.48)can be used to find the weighting coefficients <strong>for</strong> the superposition, to give as thefull solutiony(x) =∞∑ 1λ nn=0∫ ba y∗ n(z)f(z) dz∫ ba y∗ n (z)y n(z)ρ(z) dz y n(x). (17.49)If we work with normalised eigenfunctions ŷ n (x), so that∫ baŷ ∗ n(z)ŷ n (z)ρ(z) dz = 1 <strong>for</strong> all n,569


EIGENFUNCTION METHODS FOR DIFFERENTIAL EQUATIONS<strong>and</strong> we assume that we may interchange the order of summation <strong>and</strong> integration,then (17.49) can be written as∫ {b ∑ ∞ [ ] } 1y(x) =ŷ n (x)ŷ ∗λn(z) f(z) dz.nan=0The quantity in braces, which is a function of x <strong>and</strong> z only, is usually writtenG(x, z), <strong>and</strong> is the Green’s function <strong>for</strong> the problem. With this notation,wherey(x) =G(x, z) =∫ baG(x, z)f(z) dz, (17.50)∞∑n=01λ nŷ n (x)ŷ ∗ n(z). (17.51)We note that G(x, z) is determined entirely by the boundary conditions <strong>and</strong> theeigenfunctions ŷ n , <strong>and</strong> hence by L itself, <strong>and</strong> that f(z) depends purely on theRHS of the inhomogeneous equation (17.44). Thus, <strong>for</strong> a given L <strong>and</strong> boundaryconditions we can establish, once <strong>and</strong> <strong>for</strong> all, a function G(x, z) that will enableus to solve the inhomogeneous equation <strong>for</strong> any RHS. From (17.51) we also notethatG(x, z) =G ∗ (z,x). (17.52)We have already met the Green’s function in the solution of second-order differentialequations in chapter 15, as the function that satisfies the equationL[G(x, z)] = δ(x − z) (<strong>and</strong> the boundary conditions). The <strong>for</strong>mulation givenabove is an alternative, though equivalent, one.◮Find an appropriate Green’s function <strong>for</strong> the equationy ′′ + 1 y = f(x),4with boundary conditions y(0) = y(π) =0.Hence,solve<strong>for</strong>(i) f(x) =sin2x <strong>and</strong> (ii)f(x) =x/2.One approach to solving this problem is to use the methods of chapter 15 <strong>and</strong> finda complementary function <strong>and</strong> particular integral. However, in order to illustrate thetechniques developed in the present chapter we will use the superposition of eigenfunctions,which, as may easily be checked, produces the same solution.The operator on the LHS of this equation is already Hermitian under the given boundaryconditions, <strong>and</strong> so we seek its eigenfunctions. These satisfy the equationy ′′ + 1 y = λy.4This equation has the familiar solution(√ ) (√ )1y(x) =A sin − λ 1x + B cos − λ x.4 4570


17.5 SUPERPOSITION OF EIGENFUNCTIONS: GREEN’S FUNCTIONS(√ )1Now, the boundary conditions require that B =0<strong>and</strong>sin − λ π =0,<strong>and</strong>so4√1− λ = n, where n =0, ±1, ±2,....4There<strong>for</strong>e, the independent eigenfunctions that satisfy the boundary conditions arey n (x) =A n sin nx,where n is any non-negative integer, <strong>and</strong> the corresponding eigenvalues are λ n = 1 − 4 n2 .The normalisation condition further requires∫ π( ) 1/2 2A 2 n sin 2 nx dx =1 ⇒ A n = .0πComparison with (17.51) shows that the appropriate Green’s function is there<strong>for</strong>e givenbyG(x, z) = 2 ∞∑ sin nx sin nz1π − .n=0 4 n2Case (i). Using (17.50), the solution with f(x) =sin2x is given byy(x) = 2 ∫ (π ∞)∑ sin nx sin nz1π 0− sin 2zdz= 2 ∞∑n=0 4 n2 πn=0Now the integral is zero unless n = 2, in which case it is∫ π0sin 2 2zdz= π 2 .∫sin nx π1− sin nz sin 2zdz.4 n2 0Thusy(x) =− 2 sin 2x ππ 15/4 2 = − 4 sin 2x15is the full solution <strong>for</strong> f(x) =sin2x. This is, of course, exactly the solution found by usingthe methods of chapter 15.Case (ii). The solution with f(x) =x/2 isgivenby∫ ()π2∞∑ sin nx sin nz zy(x) =10 π − n=0 4 n2 2 dz = 1 ∞∑πn=0The integral may be evaluated by integrating by parts. For n ≠0,∫ [π∫z cos nzπcos nzz sin nz dz = − + dznn0=−π cos nπn] π00[ ] π sin nz+n 2 0∫sin nx π1− z sin nz dz.4 n2 0= − π(−1)n .nFor n = 0 the integral is zero, <strong>and</strong> thus∞∑y(x) = (−1) n+1 sin nxn ( 1− n2) ,n=14is the full solution <strong>for</strong> f(x) =x/2. Using the methods of subsection 15.1.2, the solutionis found to be y(x) =2x − 2π sin(x/2), which may be shown to be equal to the abovesolution by exp<strong>and</strong>ing 2x − 2π sin(x/2) as a Fourier sine series. ◭571


EIGENFUNCTION METHODS FOR DIFFERENTIAL EQUATIONS17.6 A useful generalisationSometimes we encounter inhomogeneous equations of a <strong>for</strong>m slightly more generalthan (17.1), given byLy(x) − µρ(x)y(x) =f(x) (17.53)<strong>for</strong> some Hermitian operator L, with y subject to the appropriate boundaryconditions <strong>and</strong> µ a given (i.e. fixed) constant. To solve this equation we exp<strong>and</strong>y(x) <strong>and</strong>f(x) in terms of the eigenfunctions y n (x) of the operator L, which satisfyLy n (x) =λ n ρ(x)y n (x).Working in terms of the normalised eigenfunctions ŷ n (x), we first exp<strong>and</strong> f(x)as follows:∞∑∫ bf(x) = ŷ n (x) ŷn(z)f(z)ρ(z) ∗ dzUsing (17.29) this becomes=f(x) =n=0∫ ba∫ ba= ρ(x)ρ(z)ρ(x)a∞∑ŷ n (x)ŷn(z)f(z) ∗ dz. (17.54)n=0∞∑ŷ n (x)ŷn(z)f(z) ∗ dzn=0∞∑ŷ n (x)n=0∫ baŷ ∗ n(z)f(z) dz. (17.55)Next, we exp<strong>and</strong> y(x)asy = ∑ ∞n=0 c nŷ n (x) <strong>and</strong> seek the coefficients c n . Substitutingthis <strong>and</strong> (17.55) into (17.53) we have∞∑∞∑∫ bρ(x) (λ n − µ)c n ŷ n (x) =ρ(x) ŷ n (x) ŷn(z)f(z) ∗ dz,n=0from which we find thatn=0n=0∫ ∞∑ ba ŷ∗ nn=0n=0(z)f(z) dzc n =.λ n − µHence the solution of (17.53) is given by∞∑∞∑∫ŷ n (x) by = c n ŷ n (x) =ŷ ∗λ n − µn(z)f(z) dz =From this we may identify the Green’s function∞∑ ŷ n (x)ŷ ∗G(x, z) =n(z)λ n − µ .an=0572a∫ ba∞∑n=0ŷ n (x)ŷn(z)∗ f(z) dz.λ n − µ


17.7 EXERCISESWe note that if µ = λ n ,i.e.ifµ equals one of the eigenvalues of L, thenG(x, z)becomes infinite <strong>and</strong> this method runs into difficulty. No solution then existsunless the RHS of (17.53) satisfies the relation∫ baŷ ∗ n(x)f(x) dx =0.If the spectrum of eigenvalues of the operator L is anywhere continuous,the orthonormality <strong>and</strong> closure relationships of the normalised eigenfunctionsbecome∫ ba∫ ∞0ŷ ∗ n(x)ŷ m (x)ρ(x) dx = δ(n − m),ŷ ∗ n(z)ŷ n (x)ρ(x) dn = δ(x − z).Repeating the above analysis we then find that the Green’s function is given byG(x, z) =∫ ∞0ŷ n (x)ŷ ∗ n(z)λ n − µdn.17.7 Exercises17.1 By considering 〈h|h〉, whereh = f + λg with λ real, prove that, <strong>for</strong> two functionsf <strong>and</strong> g,〈f|f〉〈g|g〉 ≥ 1 [〈f|g〉 + 4 〈g|f〉]2 .The function y(x) is real <strong>and</strong> positive <strong>for</strong> all x. Its Fourier cosine trans<strong>for</strong>m ỹ c (k)is defined by∫ ∞ỹ c (k) = y(x)cos(kx) dx,−∞<strong>and</strong> it is given that ỹ c (0) = 1. Prove thatỹ c (2k) ≥ 2[ỹ c (k)] 2 − 1.17.2 Write the homogeneous Sturm-Liouville eigenvalue equation <strong>for</strong> which y(a) =y(b) =0asL(y; λ) ≡ (py ′ ) ′ + qy + λρy =0,where p(x),q(x) <strong>and</strong>ρ(x) are continuously differentiable functions. Show that ifz(x) <strong>and</strong>F(x) satisfyL(z; λ) =F(x), with z(a) =z(b) =0,then∫ bay(x)F(x) dx =0.Demonstrate the validity of this general result by direct calculation <strong>for</strong> thespecific case in which p(x) =ρ(x) =1,q(x) =0,a = −1, b =1<strong>and</strong>z(x) =1− x 2 .17.3 Consider the real eigenfunctions y n (x) of a Sturm–Liouville equation,(py ′ ) ′ + qy + λρy =0, a ≤ x ≤ b,in which p(x), q(x) <strong>and</strong>ρ(x) are continuously differentiable real functions <strong>and</strong>p(x) does not change sign in a ≤ x ≤ b. Takep(x) as positive throughout the573


EIGENFUNCTION METHODS FOR DIFFERENTIAL EQUATIONSinterval, if necessary by changing the signs of all eigenvalues. For a ≤ x 1 ≤ x 2 ≤ b,establish the identity(λ n − λ m )∫ x2x 1ρy n y m dx = [ ]y n py m ′ − y m py n′ x2.x 1Deduce that if λ n >λ m then y n (x) must change sign between two successive zerosof y m (x).[ The reader may find it helpful to illustrate this result by sketching the first feweigenfunctions of the system y ′′ + λy =0,withy(0) = y(π) = 0, <strong>and</strong> the Legendrepolynomials P n (z) <strong>for</strong>n =2, 3, 4, 5. ]17.4 Show that the equationy ′′ + aδ(x)y + λy =0,with y(±π) =0<strong>and</strong>a real, has a set of eigenvalues λ satisfyingtan(π √ λ)= 2√ λa .Investigate the conditions under which negative eigenvalues, λ = −µ 2 ,withµreal, are possible.17.5 Use the properties of Legendre polynomials to carry out the following exercises.(a) Find the solution of (1 − x 2 )y ′′ − 2xy ′ + by = f(x), valid in the range−1 ≤ x ≤ 1 <strong>and</strong> finite at x = 0, in terms of Legendre polynomials.(b) If b =14<strong>and</strong>f(x) =5x 3 , find the explicit solution <strong>and</strong> verify it by directsubstitution.[ The first six Legendre polynomials are listed in Subsection 18.1.1. ]17.6 Starting from the linearly independent functions 1, x, x 2 , x 3 , ... , in the range0 ≤ x


17.7 EXERCISESwhere κ is a constant <strong>and</strong>{x 0 ≤ x ≤ π/2,f(x) =π − x π/2


EIGENFUNCTION METHODS FOR DIFFERENTIAL EQUATIONS17.14 Express the solution of Poisson’s equation in electrostatics,∇ 2 φ(r) =−ρ(r)/ɛ 0 ,where ρ is the non-zero charge density over a finite part of space, in the <strong>for</strong>m ofan integral <strong>and</strong> hence identify the Green’s function <strong>for</strong> the ∇ 2 operator.17.15 In the quantum-mechanical study of the scattering of a particle by a potential,a Born-approximation solution can be obtained in terms of a function y(r) thatsatisfies an equation of the <strong>for</strong>m(−∇ 2 − K 2 )y(r) =F(r).Assuming that y k (r) =(2π) −3/2 exp(ik·r) is a suitably normalised eigenfunction of−∇ 2 corresponding to eigenvalue k 2 , find a suitable Green’s function G K (r, r ′ ). Bytaking the direction of the vector r − r ′ as the polar axis <strong>for</strong> a k-space integration,show that G K (r, r ′ ) can be reduced to14π 2 |r − r ′ |∫ ∞−∞w sin wdw,w 2 − w02where w 0 = K|r − r ′ |.[ This integral can be evaluated using a contour integration (chapter 24) to give(4π|r − r ′ |) −1 exp(iK|r − r ′ |). ]17.8 Hints <strong>and</strong> answers17.1 Express the condition 〈h|h〉 ≥0 as a quadratic equation in λ <strong>and</strong> then apply the∫condition <strong>for</strong> no real roots, noting that 〈f|g〉 + 〈g|f〉 is real. To put a limit ony cos 2 kx dx, setf = y 1/2 cos kx <strong>and</strong> g = y 1/2 in the inequality.17.3 Follow an argument similar to that used <strong>for</strong> proving the reality of the eigenvalues,but integrate from x 1 to x 2 , rather than from a to b. Takex 1 <strong>and</strong> x 2 as twosuccessive zeros of y m (x) <strong>and</strong> note that, if the sign of y m is α then the sign of y m(x ′ 1 )is α whilst that of y m(x ′ 2 )is−α. Now assume that y n (x) does not change sign inthe interval <strong>and</strong> has a constant sign β; show that this leads to a contradictionbetween the signs of the two sides of the identity.17.5 (a) y = ∑ a n P n (x) witha n = n +1/2 ∫ 1f(z)P n (z) dz;b − n(n +1) −1(b) 5x 3 =2P 3 (x)+3P 1 (x), giving a 1 =1/4 <strong>and</strong>a 3 = 1, leading to y =5(2x 3 −x)/4.17.7 (a) No, ∫ gf ∗′ dx ≠ 0; (b) yes; (c) no, i ∫ f ∗ gdx ≠0;(d)yes.17.9 The normalised eigenfunctions are (2/π) 1/2 sin nx, withn an integer.y(x) =(4/π) ∑ n odd [(−1)(n−1)/2 sin nx]/[n 2 (κ − n 2 )].17.11 λ n =(n +1/2) 2 π 2 ,n=0, 1, 2,... .(a) Since y n (1)y m(1) ′ ≠ 0, the Sturm–Liouville boundary conditions are not satisfied<strong>and</strong> the appropriate weight function has to be justified by inspection. Thenormalised eigenfunctions are √ 2e −x/2 sin[(n +1/2)πx], with ρ(x) =e x .(b) y(x) =(−2/π 3 ) ∑ ∞n=0 e−x/2 sin[(n +1/2)πx]/(n +1/2) 3 .17.13 y n (x) = √ 2x −1/2 sin(nπ ln x) withλ n = −n 2 π 2 ;{ ∫ −(nπ)−2 e√a n =1 2x −1 sin(nπ ln x) dx = − √ 8(nπ) −3 <strong>for</strong> n odd,0 <strong>for</strong> n even.17.15 Use the <strong>for</strong>m of Green’s function that is the integral over all eigenvalues of the‘outer product’ of two eigenfunctions corresponding to the same eigenvalue, butwith arguments r <strong>and</strong> r ′ .576


18Special functionsIn the previous two chapters, we introduced the most important second-orderlinear ODEs in physics <strong>and</strong> engineering, listing their regular <strong>and</strong> irregular singularpoints in table 16.1 <strong>and</strong> their Sturm–Liouville <strong>for</strong>ms in table 17.1. Theseequations occur with such frequency that solutions to them, which obey particularcommonly occurring boundary conditions, have been extensively studied <strong>and</strong>given special names. In this chapter, we discuss these so-called ‘special functions’<strong>and</strong> their properties. In addition, we also discuss some special functions that arenot derived from solutions of important second-order ODEs, namely the gammafunction <strong>and</strong> related functions. These convenient functions appear in a numberof contexts, <strong>and</strong> so in section 18.12 we gather together some of their properties,with a minimum of <strong>for</strong>mal proofs.18.1 Legendre functionsLegendre’s differential equation has the <strong>for</strong>m(1 − x 2 )y ′′ − 2xy ′ + l(l +1)y =0, (18.1)<strong>and</strong> has three regular singular points, at x = −1, 1, ∞. It occurs in numerousphysical applications <strong>and</strong> particularly in problems with axial symmetry thatinvolve the ∇ 2 operator, when they are expressed in spherical polar coordinates.In normal usage the variable x in Legendre’s equation is the cosine of the polarangle in spherical polars, <strong>and</strong> thus −1 ≤ x ≤ 1. The parameter l is a given realnumber, <strong>and</strong> any solution of (18.1) is called a Legendre function.In subsection 16.1.1, we showed that x = 0 is an ordinary point of (18.1), <strong>and</strong> sowe expect to find two linearly independent solutions of the <strong>for</strong>m y = ∑ ∞n=0 a nx n .Substituting, we find∞∑ [n(n − 1)an x n−2 − n(n − 1)a n x n − 2na n x n + l(l +1)a n x n] =0,n=0577


SPECIAL FUNCTIONSwhich on collecting terms gives∞∑{(n +2)(n +1)a n+2 − [n(n +1)− l(l +1)]a n } x n =0.n=0The recurrence relation is there<strong>for</strong>e[n(n +1)− l(l +1)]a n+2 = a n , (18.2)(n +1)(n +2)<strong>for</strong> n =0, 1, 2,... . If we choose a 0 = 1 <strong>and</strong> a 1 = 0 then we obtain the solutiony 1 (x) =1− l(l +1) x2+(l − 2)l(l +1)(l +3)x4 − ··· , (18.3)2! 4!whereas on choosing a 0 = 0 <strong>and</strong> a 1 = 1 we find a second solutiony 2 (x) =x − (l − 1)(l +2) x3+(l − 3)(l − 1)(l +2)(l +4)x5 − ··· . (18.4)3! 5!By applying the ratio test to these series (see subsection 4.3.2), we find that bothseries converge <strong>for</strong> |x| < 1, <strong>and</strong> so their radius of convergence is unity, which(as expected) is the distance to the nearest singular point of the equation. Since(18.3) contains only even powers of x <strong>and</strong> (18.4) contains only odd powers, thesetwo solutions cannot be proportional to one another, <strong>and</strong> are there<strong>for</strong>e linearlyindependent. Hence, the general solution to (18.1) <strong>for</strong> |x| < 1isy(x) =c 1 y 1 (x)+c 2 y 2 (x).18.1.1 Legendre functions <strong>for</strong> integer lIn many physical applications the parameter l in Legendre’s equation (18.1) isan integer, i.e. l =0, 1, 2,... . In this case, the recurrence relation (18.2) gives[l(l +1)− l(l +1)]a l+2 = a l =0,(l +1)(l +2)i.e. the series terminates <strong>and</strong> we obtain a polynomial solution of order l. Inparticular, if l is even, then y 1 (x) in (18.3) reduces to a polynomial, whereas if l isoddthesameistrueofy 2 (x) in (18.4). These solutions (suitably normalised) arecalled the Legendre polynomials of order l; they are written P l (x) <strong>and</strong> are valid<strong>for</strong> all finite x. It is conventional to normalise P l (x) in such a way that P l (1) = 1,<strong>and</strong> as a consequence P l (−1) = (−1) l . The first few Legendre polynomials areeasily constructed <strong>and</strong> are given byP 0 (x) =1,P 2 (x) = 1 2 (3x2 − 1),P 4 (x) = 1 8 (35x4 − 30x 2 +3),P 1 (x) =x,P 3 (x) = 1 2 (5x3 − 3x),P 5 (x) = 1 8 (63x5 − 70x 3 +15x).578


18.1 LEGENDRE FUNCTIONS2P 01P 1−1P 2P 3−0.5 0.51x−1−2Figure 18.1The first four Legendre polynomials.The first four Legendre polynomials are plotted in figure 18.1.Although, according to whether l is an even or odd integer, respectively, eithery 1 (x) in (18.3) or y 2 (x) in (18.4) terminates to give a multiple of the correspondingLegendre polynomial P l (x), the other series in each case does not terminate <strong>and</strong>there<strong>for</strong>e converges only <strong>for</strong> |x| < 1. According to whether l is even or odd, wedefine Legendre functions of the second kind as Q l (x) =α l y 2 (x)orQ l (x) =β l y 1 (x),respectively, where the constants α l <strong>and</strong> β l are conventionally taken to have thevaluesα l = (−1)l/2 2 l [(l/2)!] 2l!β l = (−1)(l+1)/2 2 l−1 {[(l − 1)/2]!} 2l!<strong>for</strong> l even, (18.5)<strong>for</strong> l odd. (18.6)These normalisation factors are chosen so that the Q l (x) obey the same recurrencerelations as the P l (x) (see subsection 18.1.2).The general solution of Legendre’s equation <strong>for</strong> integer l is there<strong>for</strong>ey(x) =c 1 P l (x)+c 2 Q l (x), (18.7)579


SPECIAL FUNCTIONSwhere P l (x) is a polynomial of order l, <strong>and</strong> so converges <strong>for</strong> all x, <strong>and</strong>Q l (x) isan infinite series that converges only <strong>for</strong> |x| < 1. §By using the Wronskian method, section 16.4, we may obtain closed <strong>for</strong>ms <strong>for</strong>the Q l (x).◮Use the Wronskian method to find a closed-<strong>for</strong>m expression <strong>for</strong> Q 0 (x).From (16.25) a second solution to Legendre’s equation (18.1), with l =0,is∫ x(∫ 1u) 2vy 2 (x) =P 0 (x)[P 0 (u)] exp 2 1 − v dv du2∫ x= exp [ − ln(1 − u 2 ) ] du∫ x( )du1+x=(1 − u 2 ) = 1 ln , (18.8)21 − xwhere in the second line we have used the fact that P 0 (x) =1.All that remains is to adjust the normalisation of this solution so that it agrees with(18.5). Exp<strong>and</strong>ing the logarithm in (18.8) as a Maclaurin series we obtainy 2 (x) =x + x33 + x55 + ··· .Comparing this with the expression <strong>for</strong> Q 0 (x), using (18.4) with l = 0 <strong>and</strong> the normalisation(18.5), we find that y 2 (x) is already correctly normalised, <strong>and</strong> so( ) 1+xQ 0 (x) = 1 ln .21 − xOf course, we might have recognised the series (18.4) <strong>for</strong> l = 0, but to do so <strong>for</strong> larger lwould prove progressively more difficult. ◭Using the above method <strong>for</strong> l = 1, we find( ) 1+xQ 1 (x) = 1 2 x ln − 1.1 − xClosed <strong>for</strong>ms <strong>for</strong> higher-order Q l (x) may now be found using the recurrencerelation (18.27) derived in the next subsection. The first few Legendre functionsof the second kind are plotted in figure 18.2.18.1.2 Properties of Legendre polynomialsAs stated earlier, when encountered in physical problems the variable x inLegendre’s equation is usually the cosine of the polar angle θ in spherical polarcoordinates, <strong>and</strong> we then require the solution y(x) to be regular at x = ±1, whichcorresponds to θ =0orθ = π. For this to occur we require the equation to havea polynomial solution, <strong>and</strong> so l must be an integer. Furthermore, we also require§ It is possible, in fact, to find a second solution in terms of an infinite series of negative powers ofx that is finite <strong>for</strong> |x| > 1 (see exercise 16.16).580


18.1 LEGENDRE FUNCTIONS1Q 00.5−1−0.5−0.5−1xQ 10.5 1Q 2Figure 18.2The first three Legendre functions of the second kind.the coefficient c 2 of the function Q l (x) in (18.7) to be zero, since Q l (x) is singularat x = ±1, with the result that the general solution is simply some multiple of therelevant Legendre polynomial P l (x). In this section we will study the propertiesof the Legendre polynomials P l (x) in some detail.Rodrigues’ <strong>for</strong>mulaAs an aid to establishing further properties of the Legendre polynomials we nowdevelop Rodrigues’ representation of these functions. Rodrigues’ <strong>for</strong>mula <strong>for</strong> theP l (x) isP l (x) = 1 d l2 l l! dx l (x2 − 1) l . (18.9)To prove that this is a representation we let u =(x 2 −1) l ,sothatu ′ =2lx(x 2 −1) l−1<strong>and</strong>(x 2 − 1)u ′ − 2lxu =0.If we differentiate this expression l + 1 times using Leibnitz’ theorem, we obtain[(x 2 − 1)u (l+2) +2x(l +1)u (l+1) + l(l +1)u (l)] − 2l [ xu (l+1) +(l +1)u (l)] =0,581


SPECIAL FUNCTIONSwhich reduces to(x 2 − 1)u (l+2) +2xu (l+1) − l(l +1)u (l) =0.Changing the sign all through, we recover Legendre’s equation (18.1) with u (l) asthe dependent variable. Since, from (18.9), l is an integer <strong>and</strong> u (l) is regular atx = ±1, we may make the identificationu (l) (x) =c l P l (x), (18.10)<strong>for</strong> some constant c l that depends on l. To establish the value of c l we notethat the only term in the expression <strong>for</strong> the lth derivative of (x 2 − 1) l thatdoes not contain a factor x 2 − 1, <strong>and</strong> there<strong>for</strong>e does not vanish at x =1,is(2x) l l!(x 2 − 1) 0 . Putting x = 1 in (18.10) <strong>and</strong> recalling that P l (1) = 1, there<strong>for</strong>eshows that c l =2 l l!, thus completing the proof of Rodrigues’ <strong>for</strong>mula (18.9).◮Use Rodrigues’ <strong>for</strong>mula to show thatI l =∫ 1−1P l (x)P l (x) dx = 22l +1 . (18.11)The result is trivially obvious <strong>for</strong> l = 0 <strong>and</strong> so we assume l ≥ 1. Then, by Rodrigues’<strong>for</strong>mula,∫1 1[ ][ ]d l (x 2 − 1) l d l (x 2 − 1) lI l =dx.2 2l (l!) 2 −1 dx l dx lRepeated integration by parts, with all boundary terms vanishing, reduces this to∫ 1I l =(−1)l (x 2 − 1) l d2l2 2l (l!) 2 −1 dx 2l (x2 − 1) l dx= (2l)! ∫ 1(1 − x 2 ) l dx.2 2l (l!) 2 −1If we writeK l =∫ 1−1(1 − x 2 ) l dx,then integration by parts (taking a factor 1 as the second part) givesK l =∫ 1Writing 2lx 2 as 2l − 2l(1 − x 2 )weobtainK l =2l∫ 1−1−12lx 2 (1 − x 2 ) l−1 dx.(1 − x 2 ) l−1 dx − 2l∫ 1−1(1 − x 2 ) l dx=2lK l−1 − 2lK l<strong>and</strong> hence the recurrence relation (2l +1)K l =2lK l−1 . We there<strong>for</strong>e findK l =2l 2l − 22l +12l − 1 ··· 23 K 0 =2 l 2 l l!l!(2l +1)! 2= 22l+1 (l!) 2(2l +1)! ,which, when substituted into the expression <strong>for</strong> I l , establishes the required result. ◭582


18.1 LEGENDRE FUNCTIONSMutual orthogonalityIn section 17.4, we noted that Legendre’s equation was of Sturm–Liouville <strong>for</strong>mwith p =1− x 2 , q =0,λ = l(l +1) <strong>and</strong> ρ = 1, <strong>and</strong> that its natural intervalwas [−1, 1]. Since the Legendre polynomials P l (x) are regular at the end-pointsx = ±1, they must be mutually orthogonal over this interval, i.e.∫ 1−1P l (x)P k (x) dx =0 ifl ≠ k. (18.12)Although this result follows from the general considerations of the previouschapter, it may also be proved directly, as shown in the following example.◮Prove directly that the Legendre polynomials P l (x) are mutually orthogonal over theinterval −1


SPECIAL FUNCTIONS◮Prove the expression (18.14) <strong>for</strong> the coefficients in the Legendre polynomial expansionof a function f(x).If we multiply (18.13) by P k (x) <strong>and</strong> integrate from x = −1 tox = 1 then we obtain∫ 1∞∑∫ 1P k (x)f(x) dx = a l P k (x)P l (x) dx−1l=0∫ 1= a k P k (x)P k (x) dx =2a k−12k +1 ,where we have used the orthogonality property (18.12) <strong>and</strong> the normalisation property(18.11). ◭Generating functionA useful device <strong>for</strong> manipulating <strong>and</strong> studying sequences of functions or quantitieslabelled by an integer variable (here, the Legendre polynomials P l (x) labelled byl) isagenerating function. The generating function has perhaps its greatest utilityin the area of probability theory (see chapter 30). However, it is also a greatconvenience in our present study.The generating function <strong>for</strong>, say, a series of functions f n (x) <strong>for</strong>n =0, 1, 2,... isa function G(x, h) containing, as well as x, a dummy variable h such that∞∑G(x, h) = f n (x)h n ,n=0i.e. f n (x) is the coefficient of h n in the expansion of G in powers of h. The utilityof the device lies in the fact that sometimes it is possible to find a closed <strong>for</strong>m<strong>for</strong> G(x, h).For our study of Legendre polynomials let us consider the functions P n (x)defined by the equation∞∑G(x, h) =(1− 2xh + h 2 ) −1/2 = P n (x)h n . (18.15)As we show below, the functions so defined are identical to the Legendre polynomials<strong>and</strong> the function (1 − 2xh + h 2 ) −1/2 is in fact the generating function <strong>for</strong>them. In the process we will also deduce several useful relationships between thevarious polynomials <strong>and</strong> their derivatives.◮Show that the functions P n (x) defined by (18.15) satisfy Legendre’s equationIn the following dP n (x)/dx will be denoted by P n. ′ Firstly, we differentiate the definingequation (18.15) with respect to x <strong>and</strong> geth(1 − 2xh + h 2 ) −3/2 = ∑ P nh ′ n . (18.16)Also, we differentiate (18.15) with respect to h to yield(x − h)(1 − 2xh + h 2 ) −3/2 = ∑ nP n h n−1 . (18.17)584−1n=0


18.1 LEGENDRE FUNCTIONSEquation (18.16) can then be written, using (18.15), ash ∑ P n h n =(1− 2xh + h 2 ) ∑ P nh ′ n ,<strong>and</strong> equating the coefficients of h n+1 we obtain the recurrence relationP n = P n+1 ′ − 2xP n ′ + P n−1. ′ (18.18)Equations (18.16) <strong>and</strong> (18.17) can be combined as(x − h) ∑ P nh ′ n = h ∑ nP n h n−1 ,from which the coefficent of h n yields a second recurrence relation,xP n ′ − P n−1 ′ = nP n ; (18.19)eliminating P n−1 ′ between (18.18) <strong>and</strong> (18.19) then gives the further result(n +1)P n = P n+1 ′ − xP n. ′ (18.20)If we now take the result (18.20) with n replaced by n − 1<strong>and</strong>addx times (18.19) to itwe obtain(1 − x 2 )P n ′ = n(P n−1 − xP n ). (18.21)Finally, differentiating both sides with respect to x <strong>and</strong> using (18.19) again, we find(1 − x 2 )P n ′′ − 2xP n ′ = n[(P n−1 ′ − xP n) ′ − P n ]= n(−nP n − P n )=−n(n +1)P n ,<strong>and</strong>sotheP n defined by (18.15) do indeed satisfy Legendre’s equation. ◭The above example shows that the functions P n (x) defined by (18.15) satisfyLegendre’s equation with l = n (an integer) <strong>and</strong>, also from (18.15), these functionsare regular at x = ±1. Thus P n must be some multiple of the nth Legendrepolynomial. It there<strong>for</strong>e remains only to verify the normalisation. This is easilydone at x =1,whenG becomesG(1,h)=[(1− h) 2 ] −1/2 =1+h + h 2 + ··· ,<strong>and</strong> we can see that all the P n so defined have P n (1) = 1 as required, <strong>and</strong> are thusidentical to the Legendre polynomials.A particular use of the generating function (18.15) is in representing the inversedistance between two points in three-dimensional space in terms of Legendrepolynomials. If two points r <strong>and</strong> r ′ are at distances r <strong>and</strong> r ′ , respectively, fromthe origin, with r ′ r,however,585


SPECIAL FUNCTIONSr <strong>and</strong> r ′ must be exchanged in (18.22) or the series would not converge. Thisresult may be used, <strong>for</strong> example, to write down the electrostatic potential at apoint r due to a charge q at the point r ′ . Thus, in the case r ′


18.2 ASSOCIATED LEGENDRE FUNCTIONS18.2 Associated Legendre functionsThe associated Legendre equation has the <strong>for</strong>m](1 − x 2 )y ′′ − 2xy ′ +[l(l +1)− m21 − x 2 y =0, (18.28)which has three regular singular points at x = −1, 1, ∞ <strong>and</strong> reduces to Legendre’sequation (18.1) when m = 0. It occurs in physical applications involving theoperator ∇ 2 , when expressed in spherical polars. In such cases, −l ≤ m ≤ l <strong>and</strong>m is restricted to integer values, which we will assume from here on. As was thecase <strong>for</strong> Legendre’s equation, in normal usage the variable x is the cosine of thepolar angle in spherical polars, <strong>and</strong> thus −1 ≤ x ≤ 1. Any solution of (18.28) iscalled an associated Legendre function.The point x = 0 is an ordinary point of (18.28), <strong>and</strong> one could obtain seriessolutions of the <strong>for</strong>m y = ∑ n=0 a nx n in the same manner as that used <strong>for</strong>Legendre’s equation. In this case, however, it is more instructive to note that ifu(x) is a solution of Legendre’s equation (18.1), theny(x) =(1− x 2 ) |m|/2 d|m| u(18.29)dx |m|is a solution of the associated equation (18.28).◮Prove that if u(x) is a solution of Legendre’s equation, then y(x) given in (18.29) is asolution of the associated equation.For simplicity, let us begin by assuming that m is non-negative. Legendre’s equation <strong>for</strong> ureads(1 − x 2 )u ′′ − 2xu ′ + l(l +1)u =0,<strong>and</strong>, on differentiating this equation m times using Leibnitz’ theorem, we obtain(1 − x 2 )v ′′ − 2x(m +1)v ′ +(l − m)(l + m +1)v =0, (18.30)where v(x) =d m u/dx m . On settingy(x) =(1− x 2 ) m/2 v(x),the derivatives v ′ <strong>and</strong> v ′′ may be written as(v ′ =(1− x 2 ) −m/2 y ′ +v ′′ =(1− x 2 ) −m/2 [y ′′ +mx )1 − x y ,22mx1 − x 2 y′ + m m(m +2)x2y + y1 − x2 (1 − x 2 ) 2Substituting these expressions into (18.30) <strong>and</strong> simplifying, we obtain](1 − x 2 )y ′′ − 2xy ′ +[l(l +1)− m2y =0,1 − x 2which shows that y is a solution of the associated Legendre equation (18.28). Finally, wenote that if m is negative, the value of m 2 is unchanged, <strong>and</strong> so a solution <strong>for</strong> positive mis also a solution <strong>for</strong> the corresponding negative value of m. ◭From the two linearly independent series solutions to Legendre’s equation given587].


SPECIAL FUNCTIONSin (18.3) <strong>and</strong> (18.4), which we now denote by u 1 (x) <strong>and</strong>u 2 (x), we may obtain twolinearly-independent series solutions, y 1 (x) <strong>and</strong>y 2 (x), to the associated equationby using (18.29). From the general discussion of the convergence of power seriesgiven in section 4.5.1, we see that both y 1 (x) <strong>and</strong>y 2 (x) will also converge <strong>for</strong>|x| < 1. Hence the general solution to (18.28) in this range is given byy(x) =c 1 y 1 (x)+c 2 y 2 (x).18.2.1 Associated Legendre functions <strong>for</strong> integer lIf l <strong>and</strong> m are both integers, as is the case in many physical applications, thenthe general solution to (18.28) is denoted byy(x) =c 1 P m l (x)+c 2 Q m l (x), (18.31)where Pl m(x) <strong>and</strong>Qm l (x) are associated Legendre functions of the first <strong>and</strong> secondkind, respectively. For non-negative values of m, these functions are related to theordinary Legendre functions <strong>for</strong> integer l byP m l (x) =(1− x 2 ) m/2 dm P ldx m ,Qm l (x) =(1− x 2 ) m/2 dm Q ldx m . (18.32)We see immediately that, as required, the associated Legendre functions reduceto the ordinary Legendre functions when m = 0. Since it is m 2 that appears inthe associated Legendre equation (18.28), the associated Legendre functions <strong>for</strong>negative m values must be proportional to the corresponding function <strong>for</strong> nonnegativem. The constant of proportionality is a matter of convention. For thePl m (x) it is usual to regard the definition (18.32) as being valid also <strong>for</strong> negative mvalues. Although differentiating a negative number of times is not defined, whenP l (x) is expressed in terms of the Rodrigues’ <strong>for</strong>mula (18.9), this problem doesnot occur <strong>for</strong> −l ≤ m ≤ l. § In this case,Pl−mm (l − m)!(x) =(−1)(l + m)! P l m (x). (18.33)◮Prove the result (18.33).From (18.32) <strong>and</strong> the Rodrigues’ <strong>for</strong>mula (18.9) <strong>for</strong> the Legendre polynomials, we havePl m (x) = 12 l l! (1 − x2 m/2 dl+m)dx l+m (x2 − 1) l ,<strong>and</strong>, without loss of generality, we may assume that m is non-negative. It is convenient to§ Some authors define Pl−m (x) =Pl m(x), <strong>and</strong> similarly <strong>for</strong> the Qm l (x), in which case m is replaced by|m| in the definitions (18.32). It should be noted that, in this case, many of the results presented inthis section also require m to be replaced by |m|.588


18.2 ASSOCIATED LEGENDRE FUNCTIONSwrite (x 2 − 1) = (x +1)(x − 1) <strong>and</strong> use Leibnitz’ theorem to evaluate the derivative, whichyieldsPl m (x) = 12 l l! (1 − ∑l+mx2 ) m/2 (l + m)! d r (x +1) l d l+m−r (x − 1) l.r!(l + m − r)! dx r dx l+m−r r=0Considering the two derivative factors in a term in the summation, we note that the firstis non-zero only <strong>for</strong> r ≤ l <strong>and</strong> the second is non-zero <strong>for</strong> l + m − r ≤ l. Combining theseconditions yields m ≤ r ≤ l. Per<strong>for</strong>ming the derivatives, we thus obtainPl m (x) = 1l∑2 l l! (1 − x2 ) m/2 (l + m)! l!(x +1) l−r l!(x − 1) r−mr!(l + m − r)! (l − r)! (r − m)!r=mm/2 l!(l + m)!l∑ (x +1) l−r+ m 2 (x − 1) r− m 2=(−1)2 l r!(l + m − r)!(l − r)!(r − m)! . (18.34)r=mRepeating the above calculation <strong>for</strong> Pl−m (x) <strong>and</strong> identifying once more those terms inthe sum that are non-zero, we findPl−m−m/2l!(l − m)!(x) =(−1)∑2 l l−mr=0(x +1) l−r− m 2 (x − 1) r+ m 2r!(l − m − r)!(l − r)!(r + m)!−m/2 l!(l − m)!l∑ (x +1) l−¯r+ m 2 (x − 1)¯r− m 2=(−1)2 l (¯r − m)!(l − ¯r)!(l + m − ¯r)!¯r! , (18.35)¯r=mwhere, in the second equality, we have rewritten the summation in terms of the new index¯r = r + m. Comparing (18.34) <strong>and</strong> (18.35), we immediately arrive at the required result(18.33). ◭Since P l (x) is a polynomial of order l, we have Pl m (x) =0<strong>for</strong>|m| >l.Fromits definition, it is clear that Pl m (x) is also a polynomial of order l if m is even,but contains the factor (1 − x 2 ) to a fractional power if m is odd. In either case,Pl m (x) is regular at x = ±1. The first few associated Legendre functions of thefirst kind are easily constructed <strong>and</strong> are given by (omitting the m = 0 cases)P 1 1 (x) =(1− x2 ) 1/2 , P 1 2 (x) =3x(1 − x2 ) 1/2 ,P 2 2 (x) =3(1− x2 ), P 1 3 (x) = 3 2 (5x2 − 1)(1 − x 2 ) 1/2 ,P 2 3 (x) =15x(1 − x2 ), P 3 3 (x) = 15(1 − x2 ) 3/2 .Finally, we note that the associated Legendre functions of the second kind Q m l (x),like Q l (x), are singular at x = ±1.18.2.2 Properties of associated Legendre functions P m l (x)When encountered in physical problems, the variable x in the associated Legendreequation (as in the ordinary Legendre equation) is usually the cosine of the polarangle θ in spherical polar coordinates, <strong>and</strong> we then require the solution y(x) tobe regular at x = ±1 (corresponding to θ =0orθ = π). For this to occur, werequire l to be an integer <strong>and</strong> the coefficient c 2 of the function Q m l (x) in (18.31)589


SPECIAL FUNCTIONSto be zero, since Q m l (x) is singular at x = ±1, with the result that the generalsolution is simply some multiple of one of the associated Legendre functions ofthe first kind, Pl m (x). We will study the further properties of these functions inthe remainder of this subsection.Mutual orthogonalityAs noted in section 17.4, the associated Legendre equation is of Sturm–Liouville<strong>for</strong>m (py) ′ + qy + λρy = 0, with p =1− x 2 , q = −m 2 /(1 − x 2 ), λ = l(l +1)<strong>and</strong> ρ = 1, <strong>and</strong> its natural interval is thus [−1, 1]. Since the associated Legendrefunctions Pl m (x) are regular at the end-points x = ±1, they must be mutuallyorthogonal over this interval <strong>for</strong> a fixed value of m, i.e.∫ 1−1P m l (x)P m k (x) dx =0 ifl ≠ k. (18.36)This result may also be proved directly in a manner similar to that used <strong>for</strong> demonstratingthe orthogonality of the Legendre polynomials P l (x) in section 18.1.2.Note that the value of m must be the same <strong>for</strong> the two associated Legendrefunctions <strong>for</strong> (18.36) to hold. The normalisation condition when l = k may beobtained using the Rodrigues’ <strong>for</strong>mula, as shown in the following example.◮Show thatI lm ≡∫ 1−1Pl m (x)Pl m (x) dx = 2 (l + m)!2l +1(l − m)! . (18.37)From the definition (18.32) <strong>and</strong> the Rodrigues’ <strong>for</strong>mula (18.9) <strong>for</strong> P l (x), we may write∫1 1][ ]I lm =[(1 − x 2 ) m dl+m (x 2 − 1) l d l+m (x 2 − 1) ldx,2 2l (l!) 2 −1dx l+m dx l+mwhere the square brackets identify the factors to be used when integrating by parts.Per<strong>for</strong>ming the integration by parts l + m times, <strong>and</strong> noting that all boundary termsvanish, we obtain∫ 1[]I lm = (−1)l+m (x 2 − 1) l dl+m(1 − x 2 ) m dl+m (x 2 − 1) ldx.2 2l (l!) 2 −1 dx l+m dx l+mUsing Leibnitz’ theorem, the second factor in the integr<strong>and</strong> may be written as[]d l+m(1 − x 2 ) m dl+m (x 2 − 1) ldx l+m dx l+m∑l+m=r=0(l + m)! d r (1 − x 2 ) m d 2l+2m−r (x 2 − 1) l.r!(l + m − r)! dx r dx 2l+2m−rConsidering the two derivative factors in a term in the summation on the RHS, wesee that the first is non-zero only <strong>for</strong> r ≤ 2m, whereas the second is non-zero only <strong>for</strong>2l +2m − r ≤ 2l. Combining these conditions, we find that the only non-zero term in thesum is that <strong>for</strong> which r =2m. Thus, we may write∫I lm = (−1)l+m (l + m)! 1(1 − x 2 ) l d2m (1 − x 2 ) m d 2l (1 − x 2 ) ldx.2 2l (l!) 2 (2m)!(l − m)!dx 2m dx 2l−1590


18.2 ASSOCIATED LEGENDRE FUNCTIONSSince d 2l (1 − x 2 ) l /dx 2l =(−1) l (2l)!, <strong>and</strong> noting that (−1) 2l+2m = 1, we have1 (2l)!(l + m)!I lm =2 2l (l!) 2 (l − m)!We have already shown in section 18.1.2 that∫ 1∫ 1−1(1 − x 2 ) l dx.K l ≡ (1 − x 2 ) l dx = 22l+1 (l!) 2−1(2l +1)! ,<strong>and</strong> so we obtain the final resultI lm = 2 (l + m)!2l +1(l − m)! . ◭The orthogonality <strong>and</strong> normalisation conditions, (18.36) <strong>and</strong> (18.37) respectively,mean that the associated Legendre functions Pl m (x), with m fixed, may beused in a similar way to the Legendre polynomials to exp<strong>and</strong> any reasonablefunction f(x) on the interval |x| < 1 in a series of the <strong>for</strong>mf(x) =∞∑a m+k Pm+k(x), m (18.38)k=0where, in this case, the coefficients are given by2l +1(l − m)!a l = f(x)Pl m (x) dx.2 (l + m)! −1We note that the series takes the <strong>for</strong>m (18.38) because Pl m (x) =0<strong>for</strong>m>l.Finally, it is worth noting that the associated Legendre functions Pl m (x) mustalso obey a second orthogonality relationship. This has to be so because one mayequally well write the associated Legendre equation (18.28) in Sturm–Liouville<strong>for</strong>m (py) ′ +qy+λρy = 0, with p =1−x 2 , q = l(l+1), λ = −m 2 <strong>and</strong> ρ =(1−x 2 ) −1 ;once again the natural interval is [−1, 1]. Since the associated Legendre functionsPl m (x) are regular at the end-points x = ±1, they must there<strong>for</strong>e be mutuallyorthogonal with respect to the weight function (1 − x 2 ) −1 over this interval <strong>for</strong> afixed value of l, i.e.∫ 1−1∫ 1P m l (x)P k l (x)(1 − x 2 ) −1 dx =0 if|m| ≠ |k|. (18.39)One may also show straight<strong>for</strong>wardly that the corresponding normalisation conditionwhen m = k is given by∫ 1Pl m (x)Pl m (x)(1 − x 2 ) −1 (l + m)!dx =−1m(l − m)! .In solving physical problems, however, the orthogonality condition (18.39) is notof any practical use.591


SPECIAL FUNCTIONSGenerating functionThe generating function <strong>for</strong> associated Legendre functions can be easily derivedby combining their definition (18.32) with the generating function <strong>for</strong> the Legendrepolynomials given in (18.15). We find that(2m)!(1 − x 2 ) m/2G(x, h) =2 m m!(1 − 2hx + h 2 ) = ∑∞ Pn+m(x)h m n . (18.40)m+1/2n=0◮Derive the expression (18.40) <strong>for</strong> the associated Legendre generating function.The generating function (18.15) <strong>for</strong> the Legendre polynomials reads∞∑P n h n =(1− 2xh + h 2 ) −1/2 .n=0Differentiating both sides of this result m times (assumimg m to be non-negative), mutliplyingthrough by (1 − x 2 ) m/2 <strong>and</strong> using the definition (18.32) of the associated Legendrefunctions, we obtain∞∑Pn m h n =(1− x 2 m/2dm)dx (1 − 2xh + m h2 ) −1/2 .n=0Per<strong>for</strong>ming the derivatives on the RHS gives∞∑Pn m h n = 1 · 3 · 5 ···(2m − 1)(1 − x2 ) m/2 h m.(1 − 2xh + h 2 ) m+1/2 n=0Dividing through by h m , re-indexing the summation on the LHS <strong>and</strong> noting that, quitegenerally,1 · 3 · 5 ···(2r − 1) = 1 · 2 · 3 ···2r2 · 4 · 6 ···2r = (2r)!2 r r! ,we obtain the final result (18.40). ◭Recurrence relationsAs one might expect, the associated Legendre functions satisfy certain recurrencerelations. Indeed, the presence of the two indices n <strong>and</strong> m means that a much widerrange of recurrence relations may be derived. Here we shall content ourselveswith quoting just four of the most useful relations:Pn m+1 2mx=(1 − x 2 ) P m 1/2 n +[m(m − 1) − n(n +1)]Pn m−1 , (18.41)(2n +1)xPn m =(n + m)Pn−1 m +(n − m +1)Pn+1, m (18.42)(2n + 1)(1 − x 2 ) 1/2 Pn m = Pn+1 m+1 − P n−1 m+1 , (18.43)2(1 − x 2 ) 1/2 (Pn m ) ′ = Pnm+1 − (n + m)(n − m +1)Pn m−1 . (18.44)We note that, by virtue of our adopted definition (18.32), these recurrence relationsare equally valid <strong>for</strong> negative <strong>and</strong> non-negative values of m. These relations may592


18.3 SPHERICAL HARMONICSbe derived in a number of ways, such as using the generating function (18.40) orby differentiation of the recurrence relations <strong>for</strong> the Legendre polynomials P l (x).◮Use the recurrence relation (2n +1)P n = P n+1 ′ − P n−1 ′ <strong>for</strong> Legendre polynomials to derivethe result (18.43).Differentiating the recurrence relation <strong>for</strong> the Legendre polynomials m times, we have(2n +1) dm P ndx = dm+1 P n+1− dm+1 P n−1.m dx m+1 dx m+1Multiplying through by (1 − x 2 ) (m+1)/2 <strong>and</strong> using the definition (18.32) immediately givesthe result (18.43). ◭18.3 Spherical harmonicsThe associated Legendre functions discussed in the previous section occur mostcommonly when obtaining solutions in spherical polar coordinates of Laplace’sequation ∇ 2 u = 0 (see section 21.3.1). In particular, one finds that, <strong>for</strong> solutionsthat are finite on the polar axis, the angular part of the solution is given byΘ(θ)Φ(φ) =P m l (cos θ)(C cos mφ + D sin mφ),where l <strong>and</strong> m are integers with −l ≤ m ≤ l. This general <strong>for</strong>m is sufficientlycommon that particular functions of θ <strong>and</strong> φ called spherical harmonics aredefined <strong>and</strong> tabulated. The spherical harmonics Yl m (θ, φ) are defined by[ ] 1/2 2l +1Yl m (θ, φ) =(−1) m (l − m)!Pl m (cos θ) exp(imφ). (18.45)4π (l + m)!Using (18.33), we note thatYl−m (θ, φ) =(−1) m [ Yl m (θ, φ) ] ∗,where the asterisk denotes complex conjugation. The first few spherical harmonicsYl mm(θ, φ) ≡ Ylare as follows:√√Y0 0 = 14π , Y 1 0 = 34πcos θ,√√Y ±1301= ∓8πsin θ exp(±iφ), Y2 = 516π (3 cos2 θ − 1),√√Y ±115±2152= ∓8πsin θ cos θ exp(±iφ), Y2=32π sin2 θ exp(±2iφ).Since they contain as their θ-dependent part the solution Plm to the associatedLegendre equation, the Ylm are mutually orthogonal when integrated from −1 to+1 over d(cos θ). Their mutual orthogonality with respect to φ (0 ≤ φ ≤ 2π) iseven more obvious. The numerical factor in (18.45) is chosen to make the Ylm an593


SPECIAL FUNCTIONSorthonormal set, i.e.∫ 1 ∫ 2π−10[Yml (θ, φ) ] ∗Ym ′l (θ, φ) dφ d(cos θ) =δ ′ ll ′δ mm ′. (18.46)In addition, the spherical harmonics <strong>for</strong>m a complete set in that any reasonablefunction (i.e. one that is likely to be met in a physical situation) of θ <strong>and</strong> φ canbe exp<strong>and</strong>ed as a sum of such functions,f(θ, φ) =∞∑l∑l=0 m=−la lm Yl m (θ, φ), (18.47)the constants a lm being given by∫ 1 ∫ 2π[a lm = Yml (θ, φ) ] ∗f(θ, φ) dφ d(cos θ). (18.48)−10This is in exact analogy with a Fourier series <strong>and</strong> is a particular example of thegeneral property of Sturm–Liouville solutions.Aside from the orthonormality condition (18.46), the most important relationshipobeyed by the Ylm is the spherical harmonic addition theorem. This readsP l (cos γ) =4π2l +1l∑m=−lY ml (θ, φ)[Y ml (θ ′ ,φ ′ )] ∗ , (18.49)where (θ, φ)<strong>and</strong>(θ ′ ,φ ′ ) denote two different directions in our spherical polar coordinatesystem that are separated by an angle γ. In general, spherical trigonometry(or vector methods) shows that these angles obey the identitycos γ =cosθ cos θ ′ +sinθ sin θ ′ cos(φ − φ ′ ). (18.50)◮Prove the spherical harmonic addition theorem (18.49).For the sake of brevity, it will be useful to denote the directions (θ, φ) <strong>and</strong>(θ ′ ,φ ′ )byΩ<strong>and</strong>Ω ′ , respectively. We will also denote the element of solid angle on the sphere bydΩ =dφ d(cos θ). We begin by deriving the <strong>for</strong>m of the closure relationship obeyed by thespherical harmonics. Using (18.47) <strong>and</strong> (18.48), <strong>and</strong> reversing the order of the summation<strong>and</strong> integration, we may write∫f(Ω) = dΩ ′ f(Ω ′ ) ∑ Ylm∗ (Ω ′ )Yl m (Ω),4πlmwhere ∑ lmis a convenient shorth<strong>and</strong> <strong>for</strong> the double summation in (18.47). Thus we maywrite the closure relationship <strong>for</strong> the spherical harmonics as∑lmY ml (Ω)Y m∗l (Ω ′ )=δ(Ω − Ω ′ ), (18.51)where δ(Ω − Ω ′ ) is a Dirac delta function with the properties that δ(Ω − Ω ′ )=0ifΩ≠Ω ′<strong>and</strong> ∫ 4π δ(Ω) dΩ =1. 594


18.4 CHEBYSHEV FUNCTIONSSince δ(Ω − Ω ′ ) can depend only on the angle γ between the two directions Ω <strong>and</strong> Ω ′ ,we may also exp<strong>and</strong> it in terms of a series of Legendre polynomials of the <strong>for</strong>mδ(Ω − Ω ′ )= ∑ lb l P l (cos γ). (18.52)From (18.14), the coefficients in this expansion are given byb l ==2l +122l +14π∫ 1δ(Ω − Ω ′ )P l (cos γ) d(cos γ)−1∫ 2π ∫ 10−1δ(Ω − Ω ′ )P l (cos γ) d(cos γ) dψ,where, in the second equality, we have introduced an additional integration over anazimuthal angle ψ about the direction Ω ′ (<strong>and</strong> γ is now the polar angle measured fromΩ ′ to Ω). Since the rest of the integr<strong>and</strong> does not depend upon ψ, this is equivalentto multiplying it by 2π/2π. However, the resulting double integral now has the <strong>for</strong>m ofa solid-angle integration over the whole sphere. Moreover, when Ω = Ω ′ , the angle γseparating the two directions is zero, <strong>and</strong> so cos γ = 1. Thus, we find2l +1b l =4π P 2l +1l(1) =4π ,<strong>and</strong> combining this expression with (18.51) <strong>and</strong> (18.52) gives∑Yl m (Ω)Ylm∗ (Ω ′ )= ∑ 2l +14π P l(cos γ). (18.53)lmlComparing this result with (18.49), we see that, to complete the proof of the additiontheorem, we now only need to show that the summations in l on either side of (18.53) canbe equated term by term.That such a procedure is valid may be shown by considering an arbitrary rigid rotationof the coordinate axes, thereby defining new spherical polar coordinates ¯Ω on the sphere.Any given spherical harmonic Yl m(¯Ω)in the new coordinates can be written as a linearcombination of the spherical harmonics Yl m (Ω) of the old coordinates, all having the samevalue of l. Thus,l∑Yl m (¯Ω) = D mm′ (Ω),m ′ =−ll Ylm′where the coefficients Dlmm′ depend on the rotation; note that in this expression Ω <strong>and</strong> ¯Ωrefer to the same direction, but expressed in the two different coordinate systems. If wechoose the polar axis of the new coordinate system to lie along the Ω ′ direction, then from(18.45), with m in that equation set equal to zero, we may write√4πl∑P l (cos γ) =2l +1 Y l 0 (¯Ω) = Cl 0m′ Ylm′ (Ω)m ′ =−l<strong>for</strong> some set of coefficients Cl 0m that depend on Ω ′ . Thus, we see that the equality (18.53)does indeed hold term by term in l, thus proving the addition theorem (18.49). ◭18.4 Chebyshev functionsChebyshev’s equation has the <strong>for</strong>m(1 − x 2 )y ′′ − xy ′ + ν 2 y =0, (18.54)595


SPECIAL FUNCTIONS<strong>and</strong> has three regular singular points, at x = −1, 1, ∞. By comparing it with(18.1), we see that the Chebyshev equation is very similar in <strong>for</strong>m to Legendre’sequation. Despite this similarity, equation (18.54) does not occur very oftenin physical problems, though its solutions are of considerable importance innumerical analysis. The parameter ν is a given real number, but in nearly allpractical applications it takes an integer value. From here on we thus assumethat ν = n, wheren is a non-negative integer. As was the case <strong>for</strong> Legendre’sequation, in normal usage the variable x is the cosine of an angle, <strong>and</strong> so−1 ≤ x ≤ 1. Any solution of (18.54) is called a Chebyshev function.The point x = 0 is an ordinary point of (18.54), <strong>and</strong> so we expect to findtwo linearly independent solutions of the <strong>for</strong>m y = ∑ ∞m=0 a mx m . One could findthe recurrence relations <strong>for</strong> the coefficients a m in a similar manner to that used<strong>for</strong> Legendre’s equation in section 18.1 (see exercise 16.15). For Chebyshev’sequation, however, it is easier <strong>and</strong> more illuminating to take a different approach.In particular, we note that, on making the substitution x =cosθ, <strong>and</strong> consequentlyd/dx =(−1/ sin θ) d/dθ, Chebyshev’s equation becomes (with ν = n)d 2 ydθ 2 + n2 y =0,which is the simple harmonic equation with solutions cos nθ <strong>and</strong> sin nθ. Thecorresponding linearly independent solutions of Chebyshev’s equation are thusgiven byT n (x) =cos(n cos −1 x) <strong>and</strong> V n (x) =sin(n cos −1 x). (18.55)It is straight<strong>for</strong>ward to show that the T n (x) arepolynomials of order n, whereasthe V n (x) arenot polynomials◮Find explicit <strong>for</strong>ms <strong>for</strong> the series expansions of T n (x) <strong>and</strong> V n (x).Writing x =cosθ, it is convenient first to <strong>for</strong>m the complex superpositionT n (x)+iV n (x) =cosnθ + i sin nθ=(cosθ + i sin θ) n(= x + i √ ) n1 − x 2 <strong>for</strong> |x| ≤1.Then, on exp<strong>and</strong>ing out the last expression using the binomial theorem, we obtainT n (x) =x n − n C 2 x n−2 (1 − x 2 )+ n C 4 x n−4 (1 − x 2 ) 2 − ··· , (18.56)V n (x) = √ 1 − x 2 [n C 1 x n−1 − n C 3 x n−3 (1 − x 2 )+ n C 5 x n−5 (1 − x 2 ) 2 − ···] , (18.57)where n C r = n!/[r!(n−r)!] is a binomial coefficient. We thus see that T n (x) is a polynomialof order n, but V n (x) is not a polynomial. ◭It is conventional to define the additional functionsW n (x) =(1− x 2 ) −1/2 T n+1 (x) <strong>and</strong> U n (x) =(1− x 2 ) −1/2 V n+1 (x).596(18.58)


18.4 CHEBYSHEV FUNCTIONS1T 0T 2T 30.5T 1−1−0.50.51−0.5−1Figure 18.3The first four Chebyshev polynomials of the first kind.From (18.56) <strong>and</strong> (18.57), we see immediately that U n (x) isapolynomial of ordern, but that W n (x) isnot a polynomial. In practice, it is usual to work entirely interms of T n (x)<strong>and</strong>U n (x), which are known, respectively, as Chebyshev polynomialsof the first <strong>and</strong> second kind. In particular, we note that the general solution toChebyshev’s equation can be written in terms of these polynomials as⎧√⎨c 1 T n (x)+c 2 1 − x2 U n−1 (x) <strong>for</strong> n =1, 2, 3,...,y(x) =⎩c 1 + c 2 sin −1 x <strong>for</strong> n =0.The n = 0 solution could also be written as d 1 + c 2 cos −1 x with d 1 = c 1 + 1 2 πc 2.The first few Chebyshev polynomials of the first kind are easily constructed<strong>and</strong> are given byT 0 (x) =1,T 1 (x) =x,T 2 (x) =2x 2 − 1, T 3 (x) =4x 3 − 3x,T 4 (x) =8x 4 − 8x 2 +1,T 5 (x) =16x 5 − 20x 3 +5x.The functions T 0 (x), T 1 (x), T 2 (x) <strong>and</strong>T 3 (x) are plotted in figure 18.3. In general,the Chebyshev polynomials T n (x) satisfy T n (−x) =(−1) n T n (x), which is easilydeduced from (18.56). Similarly, it is straight<strong>for</strong>ward to deduce the following597


SPECIAL FUNCTIONS4U 2U 32U 0U 1−1 −0.5 0.5 1−2−4Figure 18.4The first four Chebyshev polynomials of the second kind.special values:T n (1) = 1, T n (−1) = (−1) n , T 2n (0) = (−1) n , T 2n+1 (0) = 0.The first few Chebyshev polynomials of the second kind are also easily found<strong>and</strong> readU 0 (x) =1,U 1 (x) =2x,U 2 (x) =4x 2 − 1, U 3 (x) =8x 3 − 4x,U 4 (x) =16x 4 − 12x 2 +1, U 5 (x) =32x 5 − 32x 3 +6x.The functions U 0 (x), U 1 (x), U 2 (x) <strong>and</strong> U 3 (x) are plotted in figure 18.4. TheChebyshev polynomials U n (x) also satisfy U n (−x) =(−1) n U n (x), which may bededuced from (18.57) <strong>and</strong> (18.58), <strong>and</strong> have the special values:U n (1) = n +1, U n (−1) = (−1) n (n +1), U 2n (0) = (−1) n , U 2n+1 (0) = 0.◮Show that the Chebyshev polynomials U n (x) satisfy the differential equation(1 − x 2 )U n ′′ (x) − 3xU n(x)+n(n ′ +2)U n (x) =0. (18.59)From (18.58), we have V n+1 =(1− x 2 ) 1/2 U n <strong>and</strong> these functions satisfy the Chebyshevequation (18.54) with ν = n +1,namely(1 − x 2 )V n+1 ′′ − xV n+1 ′ +(n +1) 2 V n+1 =0. (18.60)598


18.4 CHEBYSHEV FUNCTIONSEvaluating the first <strong>and</strong> second derivatives of V n+1 ,weobtainV ′ n+1 =(1− x 2 ) 1/2 U ′ n − x(1 − x 2 ) −1/2 U nV n+1 ′′ =(1− x 2 ) 1/2 U n ′′ − 2x(1 − x 2 ) −1/2 U n ′ − (1 − x 2 ) −1/2 U n − x 2 (1 − x 2 ) −3/2 U n .Substituting these expressions into (18.60) <strong>and</strong> dividing through by (1 − x 2 ) 1/2 , we find(1 − x 2 )U n ′′ − 3xU n ′ − U n +(n +1) 2 U n =0,which immediately simplifies to give the required result (18.59). ◭18.4.1 Properties of Chebyshev polynomialsThe Chebyshev polynomials T n (x) <strong>and</strong>U n (x) have their principal applicationsin numerical analysis. Their use in representing other functions over the range|x| < 1 plays an important role in numerical integration; Gauss–Chebyshevintegration is of particular value <strong>for</strong> the accurate evaluation of integrals whoseintegr<strong>and</strong>s contain factors (1 − x 2 ) ±1/2 . It is there<strong>for</strong>e worthwhile outlining someof their main properties.Rodrigues’ <strong>for</strong>mulaThe Chebyshev polynomials T n (x) <strong>and</strong>U n (x) may be expressed in terms of aRodrigues’ <strong>for</strong>mula, in a similar way to that used <strong>for</strong> the Legendre polynomialsdiscussed in section 18.1.2. For the Chebyshev polynomials, we haveT n (x) = (−1)n√ π(1 − x 2 ) 1/2 d n2 n (n − 1 2 )! dx n (1 − x2 ) n− 1 2 ,(−1) n√ π(n +1) d nU n (x) =2 n+1 (n + 1 2 )!(1 − x2 ) 1/2 dx n (1 − x2 ) n+ 1 2 .These Rodrigues’ <strong>for</strong>mulae may be proved in an analogous manner to that usedin section 18.1.2 when establishing the corresponding expression <strong>for</strong> the Legendrepolynomials.Mutual orthogonalityIn section 17.4, we noted that Chebyshev’s equation could be put into Sturm–Liouville <strong>for</strong>m with p =(1− x 2 ) 1/2 , q =0,λ = n 2 <strong>and</strong> ρ =(1− x 2 ) −1/2 , <strong>and</strong> itsnatural interval is thus [−1, 1]. Since the Chebyshev polynomials of the first kind,T n (x), are solutions of the Chebyshev equation <strong>and</strong> are regular at the end-pointsx = ±1, they must be mutually orthogonal over this interval with respect to theweight function ρ =(1− x 2 ) −1/2 ,i.e.∫ 1−1T n (x)T m (x)(1 − x 2 ) −1/2 dx =0 ifn ≠ m. (18.61)599


SPECIAL FUNCTIONSThe normalisation, when m = n, is easily found by making the substitutionx =cosθ <strong>and</strong> using (18.55). We immediately obtain∫ 1−1T n (x)T n (x)(1 − x 2 ) −1/2 dx ={π <strong>for</strong> n =0,π/2 <strong>for</strong> n =1, 2, 3,... .(18.62)The orthogonality <strong>and</strong> normalisation conditions mean that any (reasonable)function f(x) can be exp<strong>and</strong>ed over the interval |x| < 1inaseriesofthe<strong>for</strong>mf(x) = 1 2 a 0 +∞∑a n T n (x),n=1where the coefficients in the expansion are given by∫ 1a n = 2 f(x)T n (x)(1 − x 2 ) −1/2 dx.π −1For the Chebyshev polynomials of the second kind, U n (x), we see from (18.58)that (1 − x 2 ) 1/2 U n (x) =V n+1 (x) satisfies Chebyshev’s equation (18.54) with ν =n + 1. Thus, the orthogonality relation <strong>for</strong> the U n (x), obtained by replacing T i (x)by V i+1 (x) in equation (18.61), reads∫ 1−1U n (x)U m (x)(1 − x 2 ) 1/2 dx =0 ifn ≠ m.The corresponding normalisation condition, when n = m, can again be found bymaking the substitution x =cosθ, as illustrated in the following example.◮Show thatI ≡∫ 1−1U n (x)U n (x)(1 − x 2 ) 1/2 dx = π 2 .From (18.58), we see thatI =∫ 1−1which, on substituting x =cosθ, givesI =∫ 0πV n+1 (x)V n+1 (x)(1 − x 2 ) −1/2 dx,sin(n +1)θ sin(n +1)θ1sin θ (− sin θ) dθ = π 2 . ◭The above orthogonality <strong>and</strong> normalisation conditions allow one to exp<strong>and</strong>any (reasonable) function in the interval |x| < 1 in a series of the <strong>for</strong>mf(x) =∞∑a n U n (x),n=0600


18.4 CHEBYSHEV FUNCTIONSin which the coefficients a n are given bya n = 2 π∫ 1−1f(x)U n (x)(1 − x 2 ) 1/2 dx.Generating functionsThe generating functions <strong>for</strong> the Chebyshev polynomials of the first <strong>and</strong> secondkinds are given, respectively, by1 − xh∞G I (x, h) =1 − 2xh + h 2 = ∑T n (x)h n , (18.63)n=0∞1G II (x, h) =1 − 2xh + h 2 = ∑U n (x)h n . (18.64)These prescriptions may be proved in a manner similar to that used in section18.1.2 <strong>for</strong> the generating function of the Legendre polynomials. For theChebyshev polynomials, however, the generating functions are of less practicaluse, since most of the useful results can be obtained more easily by takingadvantage of the trigonometric <strong>for</strong>ms (18.55), as illustrated below.Recurrence relationsThere exist many useful recurrence relationships <strong>for</strong> the Chebyshev polynomialsT n (x) <strong>and</strong>U n (x). They are most easily derived by setting x =cosθ <strong>and</strong> using(18.55) <strong>and</strong> (18.58) to writeT n (x) =T n (cos θ) =cosnθ, (18.65)sin(n +1)θU n (x) =U n (cos θ) = . (18.66)sin θOne may then use st<strong>and</strong>ard <strong>for</strong>mulae <strong>for</strong> the trigonometric functions to derivea wide variety of recurrence relations. Of particular use are the trigonometricidentitiesn=0cos(n ± 1)θ =cosnθ cos θ ∓ sin nθ sin θ, (18.67)sin(n ± 1)θ =sinnθ cos θ ± cos nθ sin θ. (18.68)◮Show that the Chebyshev polynomials satisfy the recurrence relationsT n+1 (x) − 2xT n (x)+T n−1 (x) =0, (18.69)U n+1 (x) − 2xU n (x)+U n−1 (x) =0. (18.70)Adding the result (18.67) with the plus sign to the corresponding result with a minus signgivescos(n +1)θ +cos(n − 1)θ =2cosnθ cos θ.601


SPECIAL FUNCTIONSUsing (18.65) <strong>and</strong> setting x =cosθ immediately gives a rearrangement of the requiredresult (18.69). Similarly, adding the plus <strong>and</strong> minus cases of result (18.68) givessin(n +1)θ +sin(n − 1)θ =2sinnθ cos θ.Dividing through on both sides by sin θ <strong>and</strong> using (18.66) yields (18.70). ◭The recurrence relations (18.69) <strong>and</strong> (18.70) are extremely useful in the practicalcomputation of Chebyshev polynomials. For example, given the values of T 0 (x)<strong>and</strong> T 1 (x) at some point x, the result (18.69) may be used iteratively to obtainthe value of any T n (x) at that point; similarly, (18.70) may be used to calculatethe value of any U n (x) at some point x, given the values of U 0 (x) <strong>and</strong>U 1 (x) atthat point.Further recurrence relations satisfied by the Chebyshev polynomials areT n (x) =U n (x) − xU n−1 (x), (18.71)(1 − x 2 )U n (x) =xT n+1 (x) − T n+2 (x), (18.72)which establish useful relationships between the two sets of polynomials T n (x)<strong>and</strong> U n (x). The relation (18.71) follows immediately from (18.68), whereas (18.72)follows from (18.67), with n replaced by n + 1, on noting that sin 2 θ =1− x 2 .Additional useful results concerning the derivatives of Chebyshev polynomialsmay be obtained from (18.65) <strong>and</strong> (18.66), as illustrated in the following example.◮Show thatT n(x) ′ =nU n−1 (x),(1 − x 2 )U n(x) ′ =xU n (x) − (n +1)T n+1 (x).These results are most easily derived from the expressions (18.65) <strong>and</strong> (18.66) by notingthat d/dx =(−1/ sin θ) d/dθ. Thus,T n(x) ′ =− 1 d(cos nθ) n sin nθ= = nU n−1 (x).sin θ dθ sin θSimilarly, we findU n(x) ′ =− 1 [d sin(n +1)θsin θ dθ sin θwhich rearranges immediately to yield the stated result. ◭]sin(n +1)θ cos θ (n +1)cos(n +1)θ= −sin 3 θsin 2 θ= xU n(x)1 − x − (n +1)T n+1(x),2 1 − x 2Bessel’s equation has the <strong>for</strong>m18.5 Bessel functionsx 2 y ′′ + xy ′ +(x 2 − ν 2 )y =0, (18.73)which has a regular singular point at x = 0 <strong>and</strong> an essential singularity at x = ∞.The parameter ν is a given number, which we may take as ≥ 0 with no loss of602


18.5 BESSEL FUNCTIONSgenerality. The equation arises from physical situations similar to those involvingLegendre’s equation but when cylindrical, rather than spherical, polar coordinatesare employed. The variable x in Bessel’s equation is usually a multiple of a radialdistance <strong>and</strong> there<strong>for</strong>e ranges from 0 to ∞.We shall seek solutions to Bessel’s equation in the <strong>for</strong>m of infinite series. Writing(18.73) in the st<strong>and</strong>ard <strong>for</strong>m used in chapter 16, we havey ′′ + 1 )x y′ +(1 − ν2x 2 y =0. (18.74)By inspection, x = 0 is a regular singular point; hence we try a solution of the<strong>for</strong>m y = x σ ∑ ∞n=0 a nx n . Substituting this into (18.74) <strong>and</strong> multiplying the resultingequation by x 2−σ ,weobtain∞∑ [(σ + n)(σ + n − 1) + (σ + n) − ν2 ] ∞∑a n x n + a n x n+2 =0,n=0which simplifies ton=0n=0∞∑ [(σ + n) 2 − ν 2] ∞∑a n x n + a n x n+2 =0.Considering the coefficient of x 0 , we obtain the indicial equationσ 2 − ν 2 =0,<strong>and</strong> so σ = ±ν. For coefficients of higher powers of x we find[(σ +1) 2 − ν 2] a 1 =0, (18.75)[(σ + n) 2 − ν 2] a n + a n−2 =0 <strong>for</strong>n ≥ 2. (18.76)n=0Substituting σ = ±ν into (18.75) <strong>and</strong> (18.76), we obtain the recurrence relations(1 ± 2ν)a 1 =0, (18.77)n(n ± 2ν)a n + a n−2 =0 <strong>for</strong>n ≥ 2. (18.78)We consider now the <strong>for</strong>m of the general solution to Bessel’s equation (18.73) <strong>for</strong>two cases: the case <strong>for</strong> which ν is not an integer <strong>and</strong> that <strong>for</strong> which it is (includingzero).18.5.1 Bessel functions <strong>for</strong> non-integer νIf ν is a non-integer then, in general, the two roots of the indicial equation,σ 1 = ν <strong>and</strong> σ 2 = −ν, will not differ by an integer, <strong>and</strong> we may obtain two linearlyindependent solutions in the <strong>for</strong>m of Frobenius series. Special considerations doarise, however, when ν = m/2 <strong>for</strong>m =1, 3, 5,...,<strong>and</strong>σ 1 − σ 2 =2ν = m is an(odd positive) integer. When this happens, we may always obtain a solution in603


SPECIAL FUNCTIONSthe <strong>for</strong>m of a Frobenius series corresponding to the larger root, σ 1 = ν = m/2,as described above. However, <strong>for</strong> the smaller root, σ 2 = −ν = −m/2, we mustdetermine whether a second Frobenius series solution is possible by examiningthe recurrence relation (18.78), which readsn(n − m)a n + a n−2 =0 <strong>for</strong>n ≥ 2.Since m is an odd positive integer in this case, we can use this recurrence relation(starting with a 0 ≠ 0) to calculate a 2 ,a 4 ,a 6 ,... in the knowledge that all theseterms will remain finite. It is possible in this case, there<strong>for</strong>e, to find a secondsolution in the <strong>for</strong>m of a Frobenius series, one that corresponds to the smallerroot σ 2 .Thus, in general, <strong>for</strong> non-integer ν we have from (18.77) <strong>and</strong> (18.78)1a n = −n(n ± 2ν) a n−2 <strong>for</strong> n =2, 4, 6,...,= 0 <strong>for</strong> n =1, 3, 5,....Setting a 0 = 1 in each case, we obtain the two solutionsy ±ν (x) =x[1 ±ν x 2−2(2 ± 2ν) + x 4···]2 × 4(2 ± 2ν)(4 ± 2ν) − .It is customary, however, to set1a 0 =2 ±ν Γ(1 ± ν) ,where Γ(x) isthegamma function, described in subsection 18.12.1; it may beregarded as the generalisation of the factorial function to non-integer <strong>and</strong>/ornegative arguments. § The two solutions of (18.73) are then written as J ν (x) <strong>and</strong>J −ν (x), where1 xJ ν (x) =Γ(ν +1)(2∞∑=n=0) ν[1 − 1(−1) nn!Γ(ν + n +1)( x2x) 2 1 1( x) 4+ − ···]ν +1(2 (ν +1)(ν +2) 2! 2) ν+2n; (18.79)replacing ν by −ν gives J −ν (x). The functions J ν (x) <strong>and</strong>J −ν (x) are called Besselfunctions of the first kind, of order ν. Since the first term of each series is afinite non-zero multiple of x ν <strong>and</strong> x −ν , respectively, if ν is not an integer thenJ ν (x) <strong>and</strong>J −ν (x) are linearly independent. This may be confirmed by calculatingthe Wronskian of these two functions. There<strong>for</strong>e, <strong>for</strong> non-integer ν the generalsolution of Bessel’s equation (18.73) is given byy(x) =c 1 J ν (x)+c 2 J −ν (x). (18.80)§ In particular, Γ(n +1)=n! <strong>for</strong>n =0, 1, 2,..., <strong>and</strong> Γ(n) is infinite if n is any integer ≤ 0.604


18.5 BESSEL FUNCTIONSWe note that Bessel functions of half-integer order are expressible in closed <strong>for</strong>min terms of trigonometric functions, as illustrated in the following example.◮Find the general solution ofx 2 y ′′ + xy ′ +(x 2 − 1 )y =0.4This is Bessel’s equation with ν =1/2, so from (18.80) the general solution is simplyy(x) =c 1 J 1/2 (x)+c 2 J −1/2 (x).However, Bessel functions of half-integral order can be expressed in terms of trigonometricfunctions. To show this, we note from (18.79) thatJ ±1/2 (x) =x ±1/2∞∑n=0(−1) n x 2n2 2n±1/2 n!Γ(1 + n ± 1 2 ) .Using the fact that Γ(x +1)=xΓ(x) <strong>and</strong>Γ( 1 2 )=√ π,wefindthat,<strong>for</strong>ν =1/2,J 1/2 (x) = ( 1 2 x)1/2Γ( 3 2 ) − ( 1 2 x)5/21!Γ( 5 2 ) + ( 1 2 x)9/22!Γ( 7 2 ) − ···= ( 1 2 x)1/2( 1 )√ π − ( 1 2 x)5/21!( 3 )( 1 )√ π + ( 1 2 x)9/22!( 5 )( 3 )( 1 )√ π − ···2 2 2 2 2 2= ( 1 2 x)1/2( 1 2 )√ πwhereas <strong>for</strong> ν = −1/2 weobtain(1 − x23! + x45! − ···)J −1/2 (x) = ( 1 2 x)−1/2Γ( 1 2 ) − ( 1 2 x)3/21!Γ( 3 2 ) + ( 1 2 x)7/22!Γ( 5 2 ) − ···= ( 1 (···)2 x)−1/2√ 1 − x2π 2! + x44! − =There<strong>for</strong>e the general solution we require isy(x) =c 1 J 1/2 (x)+c 2 J −1/2 (x) =c 1√2πx sin x + c 2= ( √ 12 x)1/2 sin x 2( 1 )√ π x = sin x,πx2√2cos x.πx√2cos x. ◭πx18.5.2 Bessel functions <strong>for</strong> integer νThe definition of the Bessel function J ν (x) given in (18.79) is, of course, valid <strong>for</strong>all values of ν, but, as we shall see, in the case of integer ν the general solution ofBessel’s equation cannot be written in the <strong>for</strong>m (18.80). Firstly, let us consider thecase ν = 0, so that the two solutions to the indicial equation are equal, <strong>and</strong> weclearly obtain only one solution in the <strong>for</strong>m of a Frobenius series. From (18.79),605


SPECIAL FUNCTIONS1.51J 0J 1J 20.52 4 6 8 10 x−0.5Figure 18.5The first three integer-order Bessel functions of the first kind.this is given byJ 0 (x) =∞∑n=0(−1) n x 2n2 2n n!Γ(1 + n)=1− x22 2 + x42 2 4 2 − x62 2 4 2 6 2 + ··· .In general, however, if ν is a positive integer then the solutions of the indicialequation differ by an integer. For the larger root, σ 1 = ν, we may find a solutionJ ν (x), <strong>for</strong> ν =1, 2, 3,..., in the <strong>for</strong>m of the Frobenius series given by (18.79).Graphs of J 0 (x), J 1 (x) <strong>and</strong>J 2 (x) are plotted in figure 18.5 <strong>for</strong> real x. Forthesmaller root, σ 2 = −ν, however, the recurrence relation (18.78) becomesn(n − m)a n + a n−2 =0 <strong>for</strong>n ≥ 2,where m =2ν is now an even positive integer, i.e. m =2, 4, 6,... . Starting witha 0 ≠ 0 we may then calculate a 2 ,a 4 ,a 6 ,..., but we see that when n = m thecoefficient a n is <strong>for</strong>mally infinite, <strong>and</strong> the method fails to produce a secondsolution in the <strong>for</strong>m of a Frobenius series.In fact, by replacing ν by −ν in the definition of J ν (x) given in (18.79), it canbe shown that, <strong>for</strong> integer ν,J −ν (x) =(−1) ν J ν (x),606


18.5 BESSEL FUNCTIONS<strong>and</strong> hence that J ν (x) <strong>and</strong>J −ν (x) are linearly dependent. So, in this case, we cannotwrite the general solution to Bessel’s equation in the <strong>for</strong>m (18.80). One there<strong>for</strong>edefines the functionY ν (x) = J ν(x)cosνπ − J −ν (x), (18.81)sin νπwhich is called a Bessel function of the second kind of order ν (or, occasionally,a Weber or Neumann function). As Bessel’s equation is linear, Y ν (x) is clearly asolution, since it is just the weighted sum of Bessel functions of the first kind.Furthermore, <strong>for</strong> non-integer ν it is clear that Y ν (x) is linearly independent ofJ ν (x). It may also be shown that the Wronskian of J ν (x) <strong>and</strong>Y ν (x) is non-zero<strong>for</strong> all values of ν. Hence J ν (x) <strong>and</strong>Y ν (x) always constitute a pair of independentsolutions.◮If n is an integer, show that Y n+1/2 (x) =(−1) n+1 J −n−1/2 (x).From (18.81), we haveY n+1/2 (x) = J n+1/2(x)cos(n + 1 )π − J 2 −n−1/2(x)sin(n + 1 )π .2If n is an integer, cos(n + 1 )π = 0 <strong>and</strong> sin(n + 1 )π 2 2 =(−1)n , <strong>and</strong> so we immediately obtainY n+1/2 (x) =(−1) n+1 J −n−1/2 (x), as required. ◭The expression (18.81) becomes an indeterminate <strong>for</strong>m 0/0 when ν is aninteger, however. This is so because <strong>for</strong> integer ν we have cos νπ =(−1) ν <strong>and</strong>J −ν (x) =(−1) ν J ν (x). Nevertheless, this indeterminate <strong>for</strong>m can be evaluated usingl’Hôpital’s rule (see chapter 4). There<strong>for</strong>e, <strong>for</strong> integer ν, weset[ ]Jµ (x)cosµπ − J −µ (x)Y ν (x) = lim, (18.82)µ→ν sin µπwhich gives a linearly independent second solution <strong>for</strong> this case. Thus, we maywrite the general solution of Bessel’s equation, valid <strong>for</strong> all ν, asy(x) =c 1 J ν (x)+c 2 Y ν (x). (18.83)The functions Y 0 (x), Y 1 (x) <strong>and</strong>Y 2 (x) are plotted in figure 18.6Finally, we note that, in some applications, it is convenient to work withcomplex linear combinations of Bessel functions of the first <strong>and</strong> second kindsgiven byH (1)ν (x) =J ν (x)+iY ν (x), H ν(2) (x) =J ν (x) − iY ν (x);these are called, respectively, Hankel functions of the first <strong>and</strong> second kind o<strong>for</strong>der ν.607


SPECIAL FUNCTIONS10.5Y 0Y 1 Y22 4 6 8 10 x−0.5−1Figure 18.6The first three integer-order Bessel functions of the second kind.18.5.3 Properties of Bessel functions J ν (x)In physical applications, we often require that the solution is regular at x =0,but, from its definition (18.81) or (18.82), it is clear that Y ν (x) is singular atthe origin, <strong>and</strong> so in such physical situations the coefficient c 2 in (18.83) mustbe set to zero; the solution is then simply some multiple of J ν (x). These Besselfunctions of the first kind have various useful properties that are worthy offurther discussion. Unless otherwise stated, the results presented in this sectionapply to Bessel functions J ν (x) of integer <strong>and</strong> non-integer order.Mutual orthogonalityIn section 17.4, we noted that Bessel’s equation (18.73) could be put into conventionalSturm–Liouville <strong>for</strong>m with p = x, q = −ν 2 /x, λ = α 2 <strong>and</strong> ρ = x,provided αx is the argument of y. From the <strong>for</strong>m of p, we see that there is nonatural interval over which one would expect the solutions of Bessel’s equationcorresponding to different eigenvalues λ (but fixed ν) to be automatically orthogonal.Nevertheless, provided the Bessel functions satisfied appropriate boundaryconditions, we would expect them to obey an orthogonality relationship oversome interval [a, b] ofthe<strong>for</strong>m∫ baxJ ν (αx)J ν (βx) dx =0 <strong>for</strong>α ≠ β. (18.84)608


18.5 BESSEL FUNCTIONSTo determine the required boundary conditions <strong>for</strong> this result to hold, let usconsider the functions f(x) =J ν (αx) <strong>and</strong>g(x) =J ν (βx), which, as will be provedbelow, respectively satisfy the equationsx 2 f ′′ + xf ′ +(α 2 x 2 − ν 2 )f =0, (18.85)x 2 g ′′ + xg ′ +(β 2 x 2 − ν 2 )g =0. (18.86)◮Show that f(x) =J ν (αx) satisfies (18.85).If f(x) =J ν (αx) <strong>and</strong> we write w = αx, thendfdx = α dJ ν(w)d 2 f<strong>and</strong>dwdx = d2 J ν (w)2 α2 .dw 2When these expressions are substituted into (18.85), its LHS becomesx 2 α 2 d2 J ν (w)dw 2+ xα dJ ν(w)dw+(α2 x 2 − ν 2 )J ν (w)= w 2 d2 J ν (w)dw 2+ w dJ ν(w)dw+(w2 − ν 2 )J ν (w).But, from Bessel’s equation itself, this final expression is equal to zero, thus verifying thatf(x) does satisfy (18.85). ◭Now multiplying (18.86) by f(x) <strong>and</strong> (18.85) by g(x) <strong>and</strong> subtracting them givesddx [x(fg′ − gf ′ )] = (α 2 − β 2 )xfg, (18.87)wherewehaveusedthefactthatddx [x(fg′ − gf ′ )] = x(fg ′′ − gf ′′ )+(fg ′ − gf ′ ).By integrating (18.87) over any given range x = a to x = b, weobtain∫ b1[bxf(x)g(x) dx =aα 2 − β 2 xf(x)g ′ (x) − xg(x)f (x)] ′ ,awhich, on setting f(x) =J ν (αx) <strong>and</strong>g(x) =J ν (βx), becomes∫ b1[] bxJ ν (αx)J ν (βx) dx =aα 2 − β 2 βxJ ν (αx)J ν(βx) ′ − αxJ ν (βx)J ν(αx)′ . a(18.88)If α ≠ β, <strong>and</strong> the interval [a, b] is such that the expression on the RHS of (18.88)equals zero, then we obtain the orthogonality condition (18.84). This happens, <strong>for</strong>example, if J ν (αx) <strong>and</strong>J ν (βx) vanish at x = a <strong>and</strong> x = b, orifJ ν(αx) ′ <strong>and</strong>J ν(βx)′vanish at x = a <strong>and</strong> x = b, or <strong>for</strong> many more general conditions. It should benoted that the boundary term is automatically zero at the point x =0,asonemight expect from the fact that the Sturm–Liouville <strong>for</strong>m of Bessel’s equationhas p(x) =x.If α = β, the RHS of (18.88) takes the indeterminant <strong>for</strong>m 0/0. This may be609


SPECIAL FUNCTIONSevaluated using l’Hôpital’s rule, or alternatively we may calculate the relevantintegral directly.◮Evaluate the integral∫ baJ 2 ν (αx)xdx.Ignoring the integration limits <strong>for</strong> the moment,∫J 2 ν (αx)xdx= 1 α 2 ∫J 2 ν (u)u du,where u = αx. Integrating by parts yields∫∫I = Jν 2 (u)udu= 1 2 u2 Jν 2 (u) −J ν (u)J ′ ν(u)u 2 du.Now Bessel’s equation (18.73) can be rearranged asu 2 J ν (u) =ν 2 J ν (u) − uJ ν(u) ′ − u 2 J ν ′′ (u),which, on substitution into the expression <strong>for</strong> I, gives∫I = 1 2 u2 Jν 2 (u) − J ν(u)[ν ′ 2 J ν (u) − uJ ν(u) ′ − u 2 J ν ′′ (u)] du= 1 2 u2 Jν 2 (u) − 1 2 ν2 Jν 2 (u)+ 1 2 u2 [J ν(u)] ′ 2 + c.Since u = αx, the required integral is given by∫ bJν 2 (αx)xdx= 1 ) ] b [(x 2 − ν2J 2a2 α 2 ν (αx)+x 2 [J ν(αx)] ′ 2 , (18.89)awhich gives the normalisation condition <strong>for</strong> Bessel functions of the first kind. ◭Since the Bessel functions J ν (x) possess the orthogonality property (18.88), wemay exp<strong>and</strong> any reasonable function f(x), i.e. one obeying the Dirichlet conditionsdiscussed in chapter 12, in the interval 0 ≤ x ≤ b as a sum of Bessel functions ofa given (non-negative) order ν,f(x) =∞∑c n J ν (α n x), (18.90)n=0provided that the α n are chosen such that J ν (α n b) = 0. The coefficients c n are thengiven by∫ b2c n =b 2 Jν+1 2 (α f(x)J ν (α n x)xdx. (18.91)nb) 0Theintervalistakentobe0≤ x ≤ b, as then one need only ensure that theappropriate boundary condition is satisfied at x = b, since the boundary conditionat x = 0 is met automatically.610


18.5 BESSEL FUNCTIONS◮Prove the expression (18.91).If we multiply (18.90) by xJ ν (α m x) <strong>and</strong> integrate from x =0tox = b then we obtain∫ b∞∑∫ bxJ ν (α m x)f(x) dx = c n xJ ν (α m x)J ν (α n x) dx0n=00∫ b= c m Jν 2 (α m x)xdx0= 1 c 2 mb 2 J ′2 ν(α m b)= 1 c 2 mb 2 Jν+1(α 2 m b),where in the last two lines we have used (18.88) with α m = α ≠ β = α n , (18.89), the factthat J ν (α m b) = 0 <strong>and</strong> (18.95), which is proved below. ◭Recurrence relationsThe recurrence relations enjoyed by Bessel functions of the first kind, J ν (x), canbe derived directly from the power series definition (18.79).◮Prove the recurrence relationddx [xν J ν (x)] = x ν J ν−1 (x). (18.92)From the power series definition (18.79) of J ν (x) weobtainddx [xν J ν (x)] = d ∞∑ (−1) n x 2ν+2ndx 2 ν+2n n!Γ(ν + n +1)=∞∑n=0n=0= x ν ∞∑(−1) n x 2ν+2n−12 ν+2n−1 n!Γ(ν + n)n=0(−1) n x (ν−1)+2n2 (ν−1)+2n n!Γ((ν − 1) + n +1) = xν J ν−1 (x). ◭It may similarly be shown thatddx [x−ν J ν (x)] = −x −ν J ν+1 (x). (18.93)From (18.92) <strong>and</strong> (18.93) the remaining recurrence relations may be derived.Exp<strong>and</strong>ing out the derivative on the LHS of (18.92) <strong>and</strong> dividing through byx ν−1 , we obtain the relationxJ ν ′ (x)+νJ ν(x) =xJ ν−1 (x). (18.94)Similarly, by exp<strong>and</strong>ing out the derivative on the LHS of (18.93), <strong>and</strong> multiplyingthrough by x ν+1 , we findxJ ν(x) ′ − νJ ν (x) =−xJ ν+1 (x). (18.95)Adding (18.94) <strong>and</strong> (18.95) <strong>and</strong> dividing through by x givesJ ν−1 (x) − J ν+1 (x) =2J ν(x). ′ (18.96)611


SPECIAL FUNCTIONSFinally, subtracting (18.95) from (18.94) <strong>and</strong> dividing by x givesJ ν−1 (x)+J ν+1 (x) = 2ν x J ν(x). (18.97)◮Given that J 1/2 (x) =(2/πx) 1/2 sin x <strong>and</strong> that J −1/2 (x) =(2/πx) 1/2 cos x, expressJ 3/2 (x)<strong>and</strong> J −3/2 (x) in terms of trigonometric functions.From (18.95) we haveJ 3/2 (x) = 12x J 1/2(x) − J ′ 1/2 (x)= 1 ( ) 1/2 ( ) 1/2 22sin x − cos x + 12x πxπx2x( ) 1/2 ( )2 1=πx x sin x − cos x .Similarly, from (18.94) we haveJ −3/2 (x) =− 12x J −1/2(x)+J ′ −1/2 (x)= − 1 ( ) 1/2 ( ) 1/2 22cos x − sin x − 12x πxπx2x( ) 1/2 2=(− 1 )πx x cos x − sin x .( ) 1/2 2sin xπx( ) 1/2 2cos xπxWe see that, by repeated use of these recurrence relations, all Bessel functions J ν (x) ofhalfintegerorder may be expressed in terms of trigonometric functions. From their definition(18.81), Bessel functions of the second kind, Y ν (x), of half-integer order can be similarlyexpressed. ◭Finally, we note that the relations (18.92) <strong>and</strong> (18.93) may be rewritten inintegral <strong>for</strong>m as∫x ν J ν−1 (x) dx = x ν J ν (x),∫x −ν J ν+1 (x) dx = −x −ν J ν (x).If ν is an integer, the recurrence relations of this section may be proved usingthe generating function <strong>for</strong> Bessel functions discussed below. It may be shownthat Bessel functions of the second kind, Y ν (x), also satisfy the recurrence relationsderived above.Generating functionThe Bessel functions J ν (x), where ν = n is an integer, can be described by agenerating function in a way similar to that discussed <strong>for</strong> Legendre polynomials612


18.5 BESSEL FUNCTIONSin subsection 18.1.2. The generating function <strong>for</strong> Bessel functions of integer orderis given by[ ( xG(x, h) =exp h − 1 )] ∞∑= J n (x)h n . (18.98)2 hn=−∞By exp<strong>and</strong>ing the exponential as a power series, it is straightfoward to verify thatthe functions J n (x) defined by (18.98) are indeed Bessel functions of the first kind,as given by (18.79).The generating function (18.98) is useful <strong>for</strong> finding, <strong>for</strong> Bessel functions ofinteger order, properties that can often be extended to the non-integer case. Inparticular, the Bessel function recurrence relations may be derived.◮Use the generating function to prove, <strong>for</strong> integer ν, the recurrence relation (18.97), i.e.J ν−1 (x)+J ν+1 (x) = 2ν x J ν(x).Differentiating G(x, h) with respect to h we obtain(1+ 1 )G(x, h) =h 2∂G(x, h)∂h= x 2which can be written using (18.98) again as(1+ 1 )∑ ∞h 2x2n=−∞J n (x)h n =∞∑nJ n (x)h n−1 ,n=−∞∞∑nJ n (x)h n−1 .n=−∞Equating coefficients of h n we obtainx2 [J n(x)+J n+2 (x)] = (n +1)J n+1 (x),which, on replacing n by ν − 1, gives the required recurrence relation. ◭Integral representationsThe generating function (18.98) is also useful <strong>for</strong> deriving integral representationsof Bessel functions of integer order.◮Show that <strong>for</strong> integer n the Bessel function J n (x) is given byJ n (x) = 1 π∫ π0cos(nθ − x sin θ) dθ. (18.99)By exp<strong>and</strong>ing out the cosine term in the integr<strong>and</strong> in (18.99) we obtain the integral∫ πI = 1 [cos(x sin θ)cosnθ + sin(x sin θ)sinnθ] dθ. (18.100)π 0Now, we may express cos(x sin θ) <strong>and</strong> sin(x sin θ) in terms of Bessel functions by settingh =expiθ in (18.98) to give[ x]∞∑exp2 (exp iθ − exp(−iθ)) =exp(ix sin θ) = J m (x)expimθ.613m=−∞


SPECIAL FUNCTIONSUsing de Moivre’s theorem, exp iθ =cosθ + i sin θ, we then obtain∞∑exp (ix sin θ) =cos(x sin θ)+i sin(x sin θ) = J m (x)(cos mθ + i sin mθ).m=−∞Equating the real <strong>and</strong> imaginary parts of this expression gives∞∑cos(x sin θ) = J m (x)cosmθ,sin(x sin θ) =m=−∞∞∑J m (x)sinmθ.m=−∞Substituting these expressions into (18.100) then yieldsI = 1 ∞∑∫ π[J m (x)cosmθ cos nθ + J m (x)sinmθ sin nθ] dθ.πm=−∞ 0However, using the orthogonality of the trigonometric functions [ see equations (12.1)–(12.3) ], we obtainI = 1 ππ 2 [J n(x)+J n (x)] = J n (x),which proves the integral representation (18.99). ◭Finally, we mention the special case of the integral representation (18.99) <strong>for</strong>n =0:J 0 (x) = 1 ∫ πcos(x sin θ) dθ = 1 ∫ 2πcos(x sin θ) dθ,π 02π 0since cos(x sin θ) repeats itself in the range θ = π to θ =2π. However, sin(x sin θ)changes sign in this range <strong>and</strong> so∫ 2π1sin(x sin θ) dθ =0.2π 0Using de Moivre’s theorem, we can there<strong>for</strong>e writeJ 0 (x) = 12π∫ 2π0exp(ix sin θ) dθ = 12π∫ 2π0exp(ix cos θ) dθ.There are in fact many other integral representations of Bessel functions; theycan be derived from those given.18.6 Spherical Bessel functionsWhen obtaining solutions of Helmholtz’ equation (∇ 2 + k 2 )u = 0 in sphericalpolar coordinates (see section 21.3.2), one finds that, <strong>for</strong> solutions that are finiteon the polar axis, the radial part R(r) of the solution must satisfy the equationr 2 R ′′ +2rR ′ +[k 2 r 2 − l(l +1)]R =0, (18.101)614


18.6 SPHERICAL BESSEL FUNCTIONSwhere l is an integer. This equation looks very much like Bessel’s equation <strong>and</strong>can in fact be reduced to it by writing R(r) =r −1/2 S(r), in which case S(r) thensatisfies[r 2 S ′′ + rS ′ + k 2 r 2 − ( ) ]l + 1 22S =0.On making the change of variable x = kr <strong>and</strong> letting y(x) =S(kr), we obtainx 2 y ′′ + xy ′ +[x 2 − (l + 1 2 )2 ]y =0,where the primes now denote d/dx. This is Bessel’s equation of order l + 1 2<strong>and</strong> has as its solutions y(x) =J l+1/2 (x) <strong>and</strong>Y l+1/2 (x). The general solution of(18.101) can there<strong>for</strong>e be writtenR(r) =r −1/2 [c 1 J l+1/2 (kr)+c 2 Y l+1/2 (kr)],where c 1 <strong>and</strong> c 2 are constants that may be determined from the boundaryconditions on the solution. In particular, <strong>for</strong> solutions that are finite at the originwe require c 2 =0.The functions x −1/2 J l+1/2 (x) <strong>and</strong>x −1/2 Y l+1/2 (x), when suitably normalised, arecalled spherical Bessel functions of the first <strong>and</strong> second kind, respectively, <strong>and</strong> aredenoted as follows:j l (x) =√ π2x J l+1/2(x), (18.102)n l (x) =√ π2x Y l+1/2(x). (18.103)For integer l, we also note that Y l+1/2 (x) =(−1) l+1 J −l−1/2 (x), as discussed insection 18.5.2. Moreover, in section 18.5.1, we noted that Bessel functions of thefirst kind, J ν (x), of half-integer order are expressible in closed <strong>for</strong>m in terms oftrigonometric functions. Thus, all spherical Bessel functions of both the first <strong>and</strong>second kinds may be expressed in such a <strong>for</strong>m. In particular, using the results ofthe worked example in section 18.5.1, we find thatj 0 (x) = sin xx , (18.104)n 0 (x) =− cos xx . (18.105)Expressions <strong>for</strong> higher-order spherical Bessel functions are most easily obtainedby using the recurrence relations <strong>for</strong> Bessel functions.615


SPECIAL FUNCTIONS◮Show that the lth spherical Bessel function is given by( ) l 1f l (x) =(−1) l x l df 0 (x), (18.106)x dxwhere f l (x) denotes either j l (x) or n l (x).The recurrence relation (18.93) <strong>for</strong> Bessel functions of the first kind readsJ ν+1 (x) =−x ν d [x −ν J ν (x) ] .dxThus, on setting ν = l + 1 <strong>and</strong> rearranging, we find2x −1/2 J l+3/2 (x) =−x l d [ x −1/2 ]J l+1/2,dx x lwhich on using (18.102) yields the recurrence relationj l+1 (x) =−x l ddx [x−l j l (x)].We now change l +1→ l <strong>and</strong> iterate this result:j l (x) =−x l−1 d= −x l−1 ddx=(−1) 2 xlx= ···dx [ x−l+1 j l−1 (x)]{x −l+1 (−1)x l−2 ddxddx{ 1xddx[x −l+2 j l−2 (x) ]}[x −l+2 j l−2 (x) ]}( ) l 1=(−1) l x l dj 0 (x).x dxThis is the expression <strong>for</strong> j l (x) as given in (18.106). One may prove the result (18.106) <strong>for</strong>n l (x) in an analogous manner by setting ν = l − 1 in the recurrence relation (18.92) <strong>for</strong>2Bessel functions of the first kind <strong>and</strong> using the relationship Y l+1/2 (x) =(−1) l+1 J −l−1/2 (x). ◭Using result (18.106) <strong>and</strong> the expressions (18.104) <strong>and</strong> (18.105), one quicklyfinds, <strong>for</strong> example,j 1 (x) = sin xx 2 − cos x3x , j 2(x) =(x 3 − 1 )sin x − 3cosxxx 2 ,n 1 (x) =− cos xx 2− sin xx , n 2(x) =−( 3x 3 − 1 x)cos x − 3sinxx 2 .Finally, we note that the orthogonality properties of the spherical Bessel functionsfollow directly from the orthogonality condition (18.88) <strong>for</strong> Bessel functions ofthe first kind.18.7 Laguerre functionsLaguerre’s equation has the <strong>for</strong>mxy ′′ +(1− x)y ′ + νy = 0; (18.107)616


18.7 LAGUERRE FUNCTIONSit has a regular singularity at x = 0 <strong>and</strong> an essential singularity at x = ∞. Theparameter ν is a given real number, although it nearly always takes an integervalue in physical applications. The Laguerre equation appears in the descriptionof the wavefunction of the hydrogen atom. Any solution of (18.107) is called aLaguerre function.Since the point x = 0 is a regular singularity, we may find at least one solutionin the <strong>for</strong>m of a Frobenius series (see section 16.3):y(x) =∞∑a m x m+σ . (18.108)m=0Substituting this series into (18.107) <strong>and</strong> dividing through by x σ−1 ,weobtain∞∑[(m + σ)(m + σ − 1) + (1 − x)(m + σ)+νx] a m x m =0.m=0(18.109)Setting x = 0, so that only the m = 0 term remains, we obtain the indicialequation σ 2 = 0, which trivially has σ = 0 as its repeated root. Thus, Laguerre’sequation has only one solution of the <strong>for</strong>m (18.108), <strong>and</strong> it, in fact, reduces toa simple power series. Substituting σ = 0 into (18.109) <strong>and</strong> dem<strong>and</strong>ing that thecoefficient of x m+1 vanishes, we obtain the recurrence relationa m+1 =m − ν(m +1) 2 a m.As mentioned above, in nearly all physical applications, the parameter ν takesinteger values. There<strong>for</strong>e, if ν = n, wheren is a non-negative integer, we see thata n+1 = a n+2 = ···= 0, <strong>and</strong> so our solution to Laguerre’s equation is a polynomialof order n. It is conventional to choose a 0 = 1, so that the solution is given byL n (x) =[x (−1)n n − n2n! 1! xn−1 + n2 (n − 1) 2]x n−2 − ···+(−1) n n! (18.110)2!n∑= (−1) m n!(m!) 2 (n − m)! xm , (18.111)m=0where L n (x) is called the nth Laguerre polynomial. We note in particular thatL n (0) = 1. The first few Laguerre polynomials are given byL 0 (x) =1, 3!L 3 (x) =−x 3 +9x 2 − 18x +6,L 1 (x) =−x +1, 4!L 4 (x) =x 4 − 16x 3 +72x 2 − 96x +24,2!L 2 (x) =x 2 − 4x +2, 5!L 5 (x) =−x 5 +25x 4 − 200x 3 + 600x 2 − 600x + 120.The functions L 0 (x), L 1 (x), L 2 (x) <strong>and</strong>L 3 (x) are plotted in figure 18.7.617


SPECIAL FUNCTIONS105L 2L 3L 0L 11 2 3 45 6 7x−5−10Figure 18.7The first four Laguerre polynomials.18.7.1 Properties of Laguerre polynomialsThe Laguerre polynomials <strong>and</strong> functions derived from them are important inthe analysis of the quantum mechanical behaviour of some physical systems. Wethere<strong>for</strong>e briefly outline their useful properties in this section.Rodrigues’ <strong>for</strong>mulaThe Laguerre polynomials can be expressed in terms of a Rodrigues’ <strong>for</strong>mulagiven byd nL n (x) = ex (x nn! dx n e −x) , (18.112)which may be proved straight<strong>for</strong>wardly by calculating the nth derivative explicitlyusing Leibnitz’ theorem <strong>and</strong> comparing the result with (18.111). This is illustratedin the following example.618


18.7 LAGUERRE FUNCTIONS◮Prove that the expression (18.112) yields the nth Laguerre polynomial.Evaluating the nth derivative in (18.112) using Leibnitz’ theorem, we findn∑L n (x) = ex n d r x n d n−r e −xC rn! dx r dx n−rr=0n∑= ex n! n!n! r!(n − r)! (n − r)! xn−r (−1) n−r e −xr=0n∑= (−1) n−r n!r!(n − r)!(n − r)! xn−r .r=0Relabelling the summation using the index m = n − r, weobtainn∑L n (x) = (−1) m n!(m!) 2 (n − m)! xm ,m=0which is precisely the expression (18.111) <strong>for</strong> the nth Laguerre polynomial. ◭Mutual orthogonalityIn section 17.4, we noted that Laguerre’s equation could be put into Sturm–Liouville <strong>for</strong>m with p = xe −x , q =0,λ = ν <strong>and</strong> ρ = e −x , <strong>and</strong> its natural intervalis thus [0, ∞]. Since the Laguerre polynomials L n (x) aresolutionsoftheequation<strong>and</strong> are regular at the end-points, they must be mutually orthogonal over thisinterval with respect to the weight function ρ = e −x ,i.e.∫ ∞0L n (x)L k (x)e −x dx =0 ifn ≠ k.This result may also be proved directly using the Rodrigues’ <strong>for</strong>mula (18.112).Indeed, the normalisation, when k = n, is most easily found using this method.◮Show thatI ≡∫ ∞0L n (x)L n (x)e −x dx =1. (18.113)Using the Rodrigues’ <strong>for</strong>mula (18.112), we may writeI = 1 ∫ ∞L n (x) dnn! dx n (xn e −x ) dx = (−1)nn!0∫ ∞0d n L ndx nxn e −x dx,where, in the second equality, we have integrated by parts n times <strong>and</strong> used the fact that theboundary terms all vanish. When d n L n /dx n is evaluated using (18.111), only the derivativeof the m = n term survives <strong>and</strong> that has the value [ (−1) n n! n!]/[(n!) 2 0!] = (−1) n . Thuswe haveI = 1 ∫ ∞x n e −x dx =1,n! 0where, in the second equality, we use the expression (18.153) defining the gamma function(see section 18.12). ◭619


SPECIAL FUNCTIONSThe above orthogonality <strong>and</strong> normalisation conditions allow us to exp<strong>and</strong> any(reasonable) function in the interval 0 ≤ x


18.8 ASSOCIATED LAGUERRE FUNCTIONSwhich trivially rearranges to give the recurrence relation (18.115).To obtain the recurrence relation (18.116), we begin by differentiating the generatingfunction (18.114) with respect to x, which yields∂G∂x = − he−xh/(1−h)= ∑ L ′(1 − h)nh n ,2<strong>and</strong> thus we have−h ∑ L n h n =(1− h) ∑ L ′ nh n .Equating coefficients of h n on each side then gives−L n−1 = L ′ n − L ′ n−1,which immediately simplifies to give (18.116). ◭18.8 Associated Laguerre functionsThe associated Laguerre equation has the <strong>for</strong>mxy ′′ +(m +1− x)y ′ + ny = 0; (18.118)it has a regular singularity at x = 0 <strong>and</strong> an essential singularity at x = ∞. Werestrict our attention to the situation in which the parameters n <strong>and</strong> m are bothnon-negative integers, as is the case in nearly all physical problems. The associatedLaguerre equation occurs most frequently in quantum-mechanical applications.Any solution of (18.118) is called an associated Laguerre function.Solutions of (18.118) <strong>for</strong> non-negative integers n <strong>and</strong> m are given by theassociated Laguerre polynomialsL m n (x) =(−1) m dmdx m L n+m(x), (18.119)where L n (x) are the ordinary Laguerre polynomials. §◮Show that the functions L m n (x) defined in (18.119) are solutions of (18.118).Since the Laguerre polynomials L n (x) are solutions of Laguerre’s equation (18.107), wehavexL ′′n+m +(1− x)L ′ n+m +(n + m)L n+m =0.Differentiating this equation m times using Leibnitz’ theorem <strong>and</strong> rearranging, we findxL (m+2)n+m+(m +1− x)L (m+1)n+m + nL (m)n+m =0.On multiplying through by (−1) m <strong>and</strong> setting L m n =(−1) m L (m)n+m, in accord with (18.119),we obtainx(L m n ) ′′ +(m +1− x)(L m n ) ′ + nL m n =0,which shows that the functions L m n are indeed solutions of (18.118). ◭§ Note that some authors define the associated Laguerre polynomials as L m n (x) =(d m /dx m )L n (x),which is thus related to our expression (18.119) by L m n (x) =(−1) m L m n+m (x).621


SPECIAL FUNCTIONSIn particular, we note that L 0 n(x) =L n (x). As discussed in the previous section,L n (x) is a polynomial of order n <strong>and</strong> so it follows that L m n (x) is also. The first fewassociated Laguerre polynomials are easily found using (18.119):L m 0 (x) =1,L m 1 (x) =−x + m +1,2!L m 2 (x) =x 2 − 2(m +2)x +(m +1)(m +2),3!L m 3 (x) =−x 3 +3(m +3)x 2 − 3(m +2)(m +3)x +(m +1)(m +2)(m +3).Indeed, in the general case, one may show straight<strong>for</strong>wardly, from the definition(18.119) <strong>and</strong> the expression (18.111) <strong>for</strong> the ordinary Laguerre polynomials, thatL m n (x) =n∑(−1) k (n + m)!k!(n − k)!(k + m)! xk . (18.120)k=018.8.1 Properties of associated Laguerre polynomialsThe properties of the associated Laguerre polynomials follow directly from thoseof the ordinary Laguerre polynomials through the definition (18.119). We shallthere<strong>for</strong>e only briefly outline the most useful results here.Rodrigues’ <strong>for</strong>mulaA Rodrigues’ <strong>for</strong>mula <strong>for</strong> the associated Laguerre polynomials is given byL m n (x) = ex x −m d nn! dx n (xn+m e −x ). (18.121)It can be proved by evaluating the nth derivative using Leibnitz’ theorem (seeexercise 18.7).Mutual orthogonalityIn section 17.4, we noted that the associated Laguerre equation could be trans<strong>for</strong>medinto a Sturm–Liouville one with p = x m+1 e −x , q =0,λ = n <strong>and</strong> ρ = x m e −x ,<strong>and</strong> its natural interval is thus [0, ∞]. Since the associated Laguerre polynomialsL m n (x) are solutions of the equation <strong>and</strong> are regular at the end-points, thosewith the same m but differing values of the eigenvalue λ = n must be mutuallyorthogonal over this interval with respect to the weight function ρ = x m e −x ,i.e.∫ ∞0L m n (x)L m k (x)x m e −x dx =0 ifn ≠ k.This result may also be proved directly using the Rodrigues’ <strong>for</strong>mula (18.121), asmay the normalisation condition when k = n.622


18.8 ASSOCIATED LAGUERRE FUNCTIONS◮Show thatI ≡∫ ∞0L m n (x)L m n (x)x m e −x dx =(n + m)!. (18.122)n!Using the Rodrigues’ <strong>for</strong>mula (18.121), we may writeI = 1 ∫ ∞∫ ∞L m n (x) dnn! 0 dx n (xn+m e −x ) dx = (−1)n d n L m nx n+m e −x dx,n! 0 dx nwhere, in the second equality, we have integrated by parts n times <strong>and</strong> used the fact thatthe boundary terms all vanish. From (18.120) we see that d n L m n /dx n =(−1) n . Thus we have∫ ∞I = 1 x n+m e −x (n + m)!dx = ,n! 0n!where, in the second equality, we use the expression (18.153) defining the gamma function(see section 18.12). ◭The above orthogonality <strong>and</strong> normalisation conditions allow us to exp<strong>and</strong> any(reasonable) function in the interval 0 ≤ x


SPECIAL FUNCTIONSwhere, in the second equality, we have exp<strong>and</strong>ed the RHS using the binomial theorem. Onequating coefficients of h n , we immediately obtainL m n (0) =(n + m)!. ◭n!m!Recurrence relationsThe various recurrence relations satisfied by the associated Laguerre polynomialsmay be derived by differentiating the generating function (18.123) with respect toeither or both of x <strong>and</strong> h, or by differentiating with respect to x the recurrencerelations obeyed by the ordinary Laguerre polynomials, discussed in section 18.7.1.Of the many recurrence relations satisfied by the associated Laguerre polynomials,two of the most useful are as follows:(n +1)L m n+1(x) =(2n + m +1− x)L m n (x) − (n + m)L m n−1(x), (18.124)x(L m n ) ′ (x) =nL m n (x) − (n + m)L m n−1(x). (18.125)For proofs of these relations the reader is referred to exercise 18.7.Hermite’s equation has the <strong>for</strong>m18.9 Hermite functionsy ′′ − 2xy ′ +2νy =0, (18.126)<strong>and</strong> has an essential singularity at x = ∞. The parameter ν is a given realnumber, although it nearly always takes an integer value in physical applications.The Hermite equation appears in the description of the wavefunction of theharmonic oscillator. Any solution of (18.126) is called a Hermite function.Since x = 0 is an ordinary point of the equation, we may find two linearlyindependent solutions in the <strong>for</strong>m of a power series (see section 16.2):∞∑y(x) = a m x m . (18.127)m=0Substituting this series into (18.107) yields∞∑[(m +1)(m +2)a m+2 +2(ν − m)a m ] x m =0.m=0Dem<strong>and</strong>ing that the coefficient of each power of x vanishes, we obtain therecurrence relation2(ν − m)a m+2 = −(m +1)(m +2) a m.As mentioned above, in nearly all physical applications, the parameter ν takesinteger values. There<strong>for</strong>e, if ν = n, wheren is a non-negative integer, we see that624


18.9 HERMITE FUNCTIONS105−1.5H 0H 1−1−0.5H 2H 3x0.5 1 1.5−5−10Figure 18.8The first four Hermite polynomials.a n+2 = a n+4 = ···= 0, <strong>and</strong> so one solution of Hermite’s equation is a polynomialof order n. For even n, it is conventional to choose a 0 =(−1) n/2 n!/(n/2)!, whereas<strong>for</strong> odd n one takes a 1 =(−1) (n−1)/2 2n!/[ 1 2(n − 1)]!. These choices allow a generalsolution to be written asH n (x) =(2x) n − n(n − 1)(2x) n−1 +[n/2]∑=m=0n(n − 1)(n − 2)(n − 3)(2x) n−4 − ···(18.128)2!(−1) m n!m!(n − 2m)! (2x)n−2m , (18.129)where H n (x) is called the nth Hermite polynomial <strong>and</strong> the notation [n/2] denotesthe integer part of n/2. We note in particular that H n (−x) =(−1) n H n (x). Thefirst few Hermite polynomials are given byH 0 (x) =1, H 3 (x) =8x 2 − 12x,H 1 (x) =2x, H 4 (x) =16x 4 − 48x 2 +12,H 2 (x) =4x 2 − 2, H 5 (x) =32x 5 − 160x 3 + 120x.The functions H 0 (x), H 1 (x), H 2 (x) <strong>and</strong>H 3 (x) are plotted in figure 18.8.625


SPECIAL FUNCTIONS18.9.1 Properties of Hermite polynomialsThe Hermite polynomials <strong>and</strong> functions derived from them are important in theanalysis of the quantum mechanical behaviour of some physical systems. Wethere<strong>for</strong>e briefly outline their useful properties in this section.Rodrigues’ <strong>for</strong>mulaThe Rodrigues’ <strong>for</strong>mula <strong>for</strong> the Hermite polynomials is given byH n (x) =(−1) n e x2dx n (e−x2 ). (18.130)This can be proved using Leibnitz’ theorem.◮Prove the Rodrigues’ <strong>for</strong>mula (18.130) <strong>for</strong> the Hermite polynomials.Letting u = e −x2 <strong>and</strong> differentiating with respect to x, we quickly find thatu ′ +2xu =0.Differentiating this equation n + 1 times using Leibnitz’ theorem then givesu (n+2) +2xu (n+1) +2(n +1)u (n) =0,which, on introducing the new variable v =(−1) n u (n) , reduces tov ′′ +2xv ′ +2(n +1)v =0. (18.131)Now letting y = e x2 v, we may write the derivatives of v asv ′ = e −x2 (y ′ − 2xy),v ′′ = e −x2 (y ′′ − 4xy ′ +4x 2 y − 2y).Substituting these expressions into (18.131), <strong>and</strong> dividing through by e −x2 , finally yieldsHermite’s equation,y ′′ − 2xy +2ny =0,thus demonstrating that y =(−1) n e x2 d n (e −x2 )/dx n is indeed a solution. Moreover, sincethis solution is clearly a polynomial of order n, it must be some multiple of H n (x). Thenormalisation is easily checked by noting that, from (18.130), the highest-order term is(2x) n , which agrees with the expression (18.128). ◭Mutual orthogonalityWe saw in section 17.4 that Hermite’s equation could be cast in Sturm–Liouville<strong>for</strong>m with p = e −x2 , q =0,λ =2n <strong>and</strong> ρ = e −x2 , <strong>and</strong> its natural interval is thus[−∞, ∞]. Since the Hermite polynomials H n (x) are solutions of the equation <strong>and</strong>are regular at the end-points, they must be mutually orthogonal over this intervalwith respect to the weight function ρ = e −x2 ,i.e.∫ ∞−∞dnH n (x)H k (x)e −x2 dx =0 ifn ≠ k.This result may also be proved directly using the Rodrigues’ <strong>for</strong>mula (18.130).Indeed, the normalisation, when k = n, is most easily found in this way.626


18.9 HERMITE FUNCTIONS◮Show thatI ≡∫ ∞−∞H n (x)H n (x)e −x2 dx =2 n n! √ π. (18.132)Using the Rodrigues’ <strong>for</strong>mula (18.130), we may write∫ ∞∫ ∞I =(−1) n H n (x) dnd n H n0 dx n (e−x2 ) dx = e −x2 dx,−∞ dx nwhere, in the second equality, we have integrated by parts n times <strong>and</strong> used the fact thatthe boundary terms all vanish. From (18.128) we see that d n H n /dx n =2 n n!. Thus we haveI =2 n n!∫ ∞−∞e −x2 dx =2 n n! √ π,where, in the second equality, we use the st<strong>and</strong>ard result <strong>for</strong> the area under a Gaussian(see section 6.4.2). ◭The above orthogonality <strong>and</strong> normalisation conditions allow any (reasonable)function in the interval −∞ ≤ x


SPECIAL FUNCTIONSDifferentiating this <strong>for</strong>m k times with respect to h gives∞∑ H n(n − k)! hn−k = ∂k G ∂k= ex2∂hk ∂h k e−(x−h)2 =(−1) k e x2 ∂k∂x k e−(x−h)2 .n=kRelabelling the summation on the LHS using the new index m = n − k, weobtain∞∑ H m+km! hm =(−1) k x2∂ke∂x k e−(x−h)2 .m=0Setting h = 0 in this equation, we findH k (x) =(−1) k e x2dx k (e−x2 ),which is the Rodrigues’ <strong>for</strong>mula (18.130) <strong>for</strong> the Hermite polynomials. ◭The generating function (18.133) is also useful <strong>for</strong> determining special valuesof the Hermite polynomials. In particular, it is straight<strong>for</strong>ward to show thatH 2n (0) = (−1) n (2n)!/n! <strong>and</strong>H 2n+1 (0) = 0.Recurrence relationsThe two most useful recurrence relations satisfied by the Hermite polynomialsare given bydkH n+1 (x) =2xH n (x) − 2nH n−1 (x), (18.134)H n(x) ′ =2nH n−1 (x). (18.135)The first relation provides a simple iterative way of evaluating the nth Hermitepolynomials at some point x = x 0 , given the values of H 0 (x) <strong>and</strong>H 1 (x) atthatpoint. For proofs of these recurrence relations, see exercise 18.5.18.10 Hypergeometric functionsThe hypergeometric equation has the <strong>for</strong>mx(1 − x)y ′′ +[c − (a + b +1)x]y ′ − aby =0, (18.136)<strong>and</strong> has three regular singular points, at x =0, 1, ∞, but no essential singularities.The parameters a, b <strong>and</strong> c are given real numbers.In our discussions of Legendre functions, associated Legendre functions <strong>and</strong>Chebyshev functions in sections 18.1, 18.2 <strong>and</strong> 18.4, respectively, it was noted thatin each case the corresponding second-order differential equation had three regularsingular points, at x = −1, 1, ∞, <strong>and</strong> no essential singularities. The hypergeometricequation can, in fact, be considered as the ‘canonical <strong>for</strong>m’ <strong>for</strong> second-orderdifferential equations with this number of singularities. It may be shown § that,§ See, <strong>for</strong> example, J. Mathews <strong>and</strong> R. L. Walker, <strong>Mathematical</strong> <strong>Methods</strong> of <strong>Physics</strong>, 2nd edn (ReadingMA: Addision–Wesley, 1971).628


18.10 HYPERGEOMETRIC FUNCTIONSby making appropriate changes of the independent <strong>and</strong> dependent variables,any second-order differential equation with three regular singularities <strong>and</strong> anordinary point at infinity can be trans<strong>for</strong>med into the hypergeometric equation(18.136) with the singularities at = −1, 1 <strong>and</strong> ∞. As we discuss below, this allowsLegendre functions, associated Legendre functions <strong>and</strong> Chebyshev functions, <strong>for</strong>example, to be written as particular cases of hypergeometric functions, which arethe solutions to (18.136).Since the point x = 0 is a regular singularity of (18.136), we may find at leastone solution in the <strong>for</strong>m of a Frobenius series (see section 16.3):y(x) =∞∑a n x n+σ . (18.137)Substituting this series into (18.136) <strong>and</strong> dividing through by x σ−1 ,weobtain∞∑n=0n=0{(1 − x)(n + σ)(n + σ − 1) + [c − (a + b +1)x](n + σ) − abx} a n x n =0.(18.138)Setting x = 0, so that only the n = 0 term remains, we obtain the indicial equationσ(σ − 1) + cσ = 0, which has the roots σ = 0 <strong>and</strong> σ =1− c. Thus, providedc is not an integer, one can obtain two linearly independent solutions of thehypergeometric equation in the <strong>for</strong>m (18.137).For σ = 0 the corresponding solution is a simple power series. Substitutingσ = 0 into (18.138) <strong>and</strong> dem<strong>and</strong>ing that the coefficient of x n vanishes, we find therecurrence relationn[(n − 1) + c]a n − [(n − 1)(a + b + n − 1) + ab]a n−1 =0,(18.139)which, on simplifying <strong>and</strong> replacing n by n + 1, yields the recurrence relation(a + n)(b + n)a n+1 =(n +1)(c + n) a n. (18.140)It is conventional to make the simple choice a 0 = 1. Thus, provided c is not anegative integer or zero, we may write the solution as follows:F(a, b, c; x) =1+ ab x a(a +1)b(b +1)+ + ···c 1! c(c +1) 2!(18.141)= Γ(c) ∞∑ Γ(a + n)Γ(b + n) x nΓ(a)Γ(b) Γ(c + n) n! , (18.142)n=0where F(a, b, c; x) is known as the hypergeometric function or hypergeometricseries, <strong>and</strong> in the second equality we have used the property (18.154) of the629x 2


SPECIAL FUNCTIONSgamma function. § It is straight<strong>for</strong>ward to show that the hypergeometric seriesconverges in the range |x| < 1. It also converges at x =1ifc>a+ b <strong>and</strong>at x = −1 ifc>a+ b − 1. We also note that F(a, b, c; x) is symmetric in theparameters a <strong>and</strong> b, i.e.F(a, b, c; x) =F(b, a, c; x).The hypergeometric function y(x) = F(a, b, c; x) is clearly not the generalsolution to the hypergeometric equation (18.136), since we must also consider thesecond root of the indicial equation. Substituting σ =1− c into (18.138) <strong>and</strong>dem<strong>and</strong>ing that the coefficient of x n vanishes, we find that we must haven(n +1− c)a n − [(n − c)(a + b + n − c)+ab]a n−1 =0,which, on comparing with (18.139) <strong>and</strong> replacing n by n + 1, yields the recurrencerelation(a − c +1+n)(b − c +1+n)a n+1 = a n .(n + 1)(2 − c + n)We see that this recurrence relation has the same <strong>for</strong>m as (18.140) if one makesthe replacements a → a − c +1,b → b − c + 1 <strong>and</strong> c → 2 − c. Thus, provided c,a−b <strong>and</strong> c−a−b are all non-integers, the general solution to the hypergeometricequation, valid <strong>for</strong> |x| < 1, may be written asy(x) =AF(a, b, c; x)+Bx 1−c F(a − c +1,b− c +1, 2 − c; x),(18.143)where A <strong>and</strong> B are arbitrary constants to be fixed by the boundary conditions onthe solution. If the solution is to be regular at x =0,onerequiresB =0.18.10.1 Properties of hypergeometric functionsSince the hypergeometric equation is so general in nature, it is not feasible topresent a comprehensive account of the hypergeometric functions. Nevertheless,we outline here some of their most important properties.Special casesAs mentioned above, the general nature of the hypergeometric equation allows usto write a large number of elementary functions in terms of the hypergeometricfunctions F(a, b, c; x). Such identifications can be made from the series expansion(18.142) directly, or by trans<strong>for</strong>mation of the hypergeometric equation into a morefamiliar equation, the solutions to which are already known. Some particularexamples of well known special cases of the hypergeometric function are asfollows:§ We note that it is also common to denote the hypergeometric function by 2 F 1 (a, b, c; x). Thisslightly odd-looking notation is meant to signify that, in the coefficient of each power of x, thereare two parameters (a <strong>and</strong> b) in the numerator <strong>and</strong> one parameter (c) in the denominator.630


18.10 HYPERGEOMETRIC FUNCTIONSF(a, b, b; x) =(1− x) −a , F( 1 2 , 1 2 , 3 2 ; x2 )=x −1 sin −1 x,F(1, 1, 2; −x) =x −1 ln(1 + x), F( 1 2 , 1, 3 2 ; −x2 )=x −1 tan −1 x,lim m→∞ =ex , F( 1 2 , 1, 3 2 ; x2 )= 1 2 x−1 ln[(1 + x)/(1 − x)],F( 1 2 , − 1 2 , 1 2 ;sin2 x)=cosx, F(m +1, −m, 1; (1 − x)/2) = P m (x),F( 1 2 ,p,p;sin2 x)=secx, F(m, −m, 1 2 ;(1− x)/2) = T m(x),where m is an integer, P m (x) isthemth Legendre polynomial <strong>and</strong> T m (x) isthemth Chebyshev polynomial of the first kind. Some of these results are proved inexercise 18.11.◮Show that F(m, −m, 1 2 ;(1− x)/2) = T m(x).Let us prove this result by trans<strong>for</strong>ming the hypergeometric equation. The <strong>for</strong>m of theresult suggests that we should make the substitution x =(1− z)/2 into (18.136), in whichcase d/dx = −2d/dz. Thus, letting u(z) =y(x) <strong>and</strong> setting a = m, b = −m <strong>and</strong> c =1/2,(18.136) becomes(1 − z)2(1 + z)(−2) 2 d2 u2On simplifying, we obtaindz 2 + [12]− z− (m − m +1)1 (−2) du − (m)(−m)u =0.2 dz(1 − z 2 ) d2 udz − z du2 dz + m2 u =0,which has the <strong>for</strong>m of Chebyshev’s equation, (18.54). This equation has u(z) =T m (z) asits power series solution, <strong>and</strong> so F(m, −m, 1 ;(1− z)/2) <strong>and</strong> T 2 m(z) are equal to within anormalisation factor. On comparing the expressions (18.141) <strong>and</strong> (18.56) at x =0,i.e.atz = 1, we see that they both have value 1. Hence, the normalisations already agree <strong>and</strong>we obtain the required result. ◭Integral representationOne of the most useful representations <strong>for</strong> the hypergeometric functions is interms of an integral, which may be derived using the properties of the gamma<strong>and</strong> beta functions discussed in section 18.12. The integral representation readsF(a, b, c; x) =Γ(c)Γ(b)Γ(c − b)∫ 1<strong>and</strong> requires c>b>0 <strong>for</strong> the integral to converge.◮Prove the result (18.144).0t b−1 (1 − t) c−b−1 (1 − tx) −a dt,From the series expansion (18.142), we haveF(a, b, c; x) =Γ(c) ∞∑ Γ(a + n)Γ(b + n) x nΓ(a)Γ(b) Γ(c + n) n!=n=0Γ(c)Γ(a)Γ(b)Γ(c − b)631∞∑n=0Γ(a + n)B(b + n, c − b) xnn! ,(18.144)


SPECIAL FUNCTIONSwhere in the second equality we have used the expression (18.165) relating the gamma <strong>and</strong>beta functions. Using the definition (18.162) of the beta function, we then findΓ(c)∞∑F(a, b, c; x) =Γ(a + n) xnΓ(a)Γ(b)Γ(c − b)n!=Γ(c)Γ(b)Γ(c − b)∫ 10n=0dt t b−1 (1 − t) c−b−1∫ 10t b+n−1 (1 − t) c−b−1 dt∞∑n=0Γ(a + n) (tx) nΓ(a) n! ,where in the second equality we have rearranged the expression <strong>and</strong> reversed the orderof integration <strong>and</strong> summation. Finally, one recognises the sum over n as being equal to(1 − tx) −a , <strong>and</strong> so we obtain the final result (18.144). ◭The integral representation may be used to prove a wide variety of properties ofthe hypergeometric functions. As a simple example, on setting x = 1 in (18.144),<strong>and</strong> using properties of the beta function discussed in section 18.12.2, one quicklyfinds that, provided c is not a negative integer or zero <strong>and</strong> c>a+ b,F(a, b, c;1)=Γ(c)Γ(c − a − b)Γ(c − a)Γ(c − b) .Relationships between hypergeometric functionsThere exist a great many relationships between hypergeometric functions withdifferent arguments. These are most easily derived by making use of the integralrepresentation (18.144) or the series <strong>for</strong>m (18.141). It is not feasible to list all therelationships here, so we simply note two useful examples, which readF(a, b, c; x) =(1− x) c−a−b F(c − a, c − b, c; x), (18.145)F ′ (a, b, c; x) = ab F(a +1,b+1,c+1;x), (18.146)cwhere the prime in the second relation denotes d/dx. The first result followsstraight<strong>for</strong>wardly from the integral representation using the substitution t =(1 − u)/(1 − ux), whereas the second result may be proved more easily from theseries expansion.In addition to the above results, one may also derive relationships betweenF(a, b, c; x) <strong>and</strong> any two of the six ‘contiguous functions’ F(a ± 1,b,c; x), F(a, b ±1,c; x) <strong>and</strong>F(a, b, c ± 1; x). These ‘contiguous relations’ serve as the recurrencerelations <strong>for</strong> the hypergeometric functions. An example of such a relationship is(c − a)F(a − 1,b,c; x)+(2a − c − ax + bx)F(a, b, c; x)+a(x − 1)F(a +1,b,c; x)=0.Repeated application of such relationships allows one to express F(a+l,b+m, c+n; x), where l,m,n are integers (with c+n not equalling a negative integer or zero),as a linear combination of F(a, b, c; x) <strong>and</strong> one of its contiguous functions.632


18.11 CONFLUENT HYPERGEOMETRIC FUNCTIONS18.11 Confluent hypergeometric functionsThe confluent hypergeometric equation has the <strong>for</strong>mxy ′′ +(c − x)y ′ − ay = 0; (18.147)it has a regular singularity at x = 0 <strong>and</strong> an essential singularity at x = ∞.This equation can be obtained by merging two of the singularities of the ordinaryhypergeometric equation (18.136). The parameters a <strong>and</strong> c are given real numbers.◮Show that setting x = z/b in the hypergeometric equation, <strong>and</strong> letting b →∞, yields theconfluent hypergeometric equation.Substituting x = z/b into (18.136), with d/dx = bd/dz, <strong>and</strong> letting u(z) =y(x), we obtain(bz 1 − z ) d 2 udu+[bc − (a + b +1)z] − abu =0,b dz2 dzwhich clearly has regular singular points at z =0,b <strong>and</strong> ∞. Ifwenowmergethelasttwosingularities by letting b →∞,weobtainzu ′′ +(c − z)u ′ − au =0,where the primes denote d/dz. Hence u(z) must satisfy the confluent hypergeometricequation. ◭In our discussion of Bessel, Laguerre <strong>and</strong> associated Laguerre functions, it wasnoted that the corresponding second-order differential equation in each case had asingle regular singular point at x = 0 <strong>and</strong> an essential singularity at x = ∞. Fromtable 16.1, we see that this is also true <strong>for</strong> the confluent hypergeometric equation.Indeed, this equation can be considered as the ‘canonical <strong>for</strong>m’ <strong>for</strong> secondorderdifferential equations with this pattern of singularities. Consequently, as wemention below, the Bessel, Laguerre <strong>and</strong> associated Laguerre functions can all bewritten in terms of the confluent hypergeometric functions, which are the solutionsof (18.147).The solutions of the confluent hypergeometric equation are obtained fromthose of the ordinary hypergeometric equation by again letting x → x/b <strong>and</strong>carrying out the limiting process b →∞. Thus, from (18.141) <strong>and</strong> (18.143), twolinearly independent solutions of (18.147) are (when c is not an integer)y 1 (x) =1+ a x a(a +1) z 2+ + ···≡ M(a, c; x), (18.148)c 1! c(c +1) 2!y 2 (x) =x 1−c M(a − c +1, 2 − c; x), (18.149)where M(a, c; x) is called the confluent hypergeometric function (or Kummerfunction). § It is worth noting, however, that y 1 (x) is singular when c =0, −1, −2,...<strong>and</strong> y 2 (x) is singular when c =2, 3, 4,... . Thus, it is conventional to take the§ We note that an alternative notation <strong>for</strong> the confluent hypergeometric function is 1 F 1 (a, c; x).633


SPECIAL FUNCTIONSsecond solution to (18.147) as a linear combination of (18.148) <strong>and</strong> (18.149) givenbyU(a, c; x) ≡π []M(a, c; x) M(a − c +1, 2 − c; x)− x1−c .sin πc Γ(a − c +1)Γ(c) Γ(a)Γ(2 − c)This has a well behaved limit as c approaches an integer.18.11.1 Properties of confluent hypergeometric functionsThe properties of confluent hypergeometric functions can be derived from thoseof ordinary hypergeometric functions by letting x → x/b <strong>and</strong> taking the limitb →∞, in the same way as both the equation <strong>and</strong> its solution were derived. Ageneral procedure of this sort is called a confluence process.Special casesThe general nature of the confluent hypergeometric equation allows one to writea large number of elementary functions in terms of the confluent hypergeometricfunctions M(a, c; x). Once again, such identifications can be made from the seriesexpansion (18.148) directly, or by trans<strong>for</strong>mation of the confluent hypergeometricequation into a more familiar equation <strong>for</strong> which the solutions are alreadyknown. Some particular examples of well known special cases of the confluenthypergeometric function are as follows:M(a, a; x) =e x ,M(1, 2; 2x) = ex sinh x,xM(−n, 1; x) =L n (x), M(−n, m +1;x) = n!m!(n + m)! Lm n (x),M(−n, 1 2 ; x2 )= (−1)n n!(2n)! H 2n(x), M(−n, 3 2 ; x2 ), = (−1)n n!2(2n +1)!M(ν + 1 2 , 2ν +1;2ix) =ν!eix ( x 2 )−ν J ν (x), M( 1 2 , 3 2 ; −x2 )=√ π2x erf(x),H 2n+1 (x),xwhere n <strong>and</strong> m are integers, L m n (x) is an associated Legendre polynomial, H n (x) isa Hermite polynomial, J ν (x) is a Bessel function <strong>and</strong> erf(x) is the error functiondiscussed in section 18.12.4.Integral representationUsing the integral representation (18.144) of the ordinary hypergeometric function,exchanging a <strong>and</strong> b <strong>and</strong> carrying out the process of confluence givesM(a, c, x) =Γ(c)Γ(a)Γ(c − a)634∫ 10e tx t a−1 (1 − t) c−a−1 dt,(18.150)


18.12 THE GAMMA FUNCTION AND RELATED FUNCTIONSwhich converges provided c>a>0.◮Prove the result (18.150).Since F(a, b, c; x) is unchanged by swapping a <strong>and</strong> b, we may write its integral representation(18.144) asΓ(c)F(a, b, c; x) =t a−1 (1 − t) c−a−1 (1 − tx) −b dt.Γ(a)Γ(c − a) 0Setting x = z/b <strong>and</strong> taking the limit b →∞,weobtainM(a, c; z) =Γ(c)Γ(a)Γ(c − a)∫ 10∫ 1t a−1 (1 − t) c−a−1 limb→∞(1 − tz bSince the limit is equal to e tz , we obtain result (18.150). ◭) −bdt.Relationships between confluent hypergeometric functionsA large number of relationships exist between confluent hypergeometric functionswith different arguments. These are straight<strong>for</strong>wardly derived using the integralrepresentation (18.150) or the series <strong>for</strong>m (18.148). Here, we simply note twouseful examples, which readM(a, c; x) =e x M(c − a, c; −x), (18.151)M ′ (a, c; x) = a M(a +1,c+1;x), (18.152)cwhere the prime in the second relation denotes d/dx. The first result followsstraight<strong>for</strong>wardly from the integral representation, <strong>and</strong> the second result may beproved from the series expansion (see exercise 18.19).In an analogous manner to that used <strong>for</strong> the ordinary hypergeometric functions,one may also derive relationships between M(a, c; x) <strong>and</strong> any two of thefour ‘contiguous functions’ M(a ± 1,c; x) <strong>and</strong>M(a, c ± 1; x). These serve as therecurrence relations <strong>for</strong> the confluent hypergeometric functions. An example ofsuch a relationship is(c − a)M(a − 1,c; x)+(2a − c + x)M(a, c; x) − aM(a +1,c; x) =0.18.12 The gamma function <strong>and</strong> related functionsMany times in this chapter, <strong>and</strong> often throughout the rest of the book, we havemade mention of the gamma function <strong>and</strong> related functions such as the beta <strong>and</strong>error functions. Although not derived as the solutions of important second-orderODEs, these convenient functions appear in a number of contexts, <strong>and</strong> so herewe gather together some of their properties. This final section should be regardedmerely as a reference containing some useful relations obeyed by these functions;a minimum of <strong>for</strong>mal proofs is given.635


SPECIAL FUNCTIONSThe gamma function Γ(n) is defined by18.12.1 The gamma functionΓ(n) =∫ ∞0x n−1 e −x dx, (18.153)which converges <strong>for</strong> n>0, where in general n is a real number. Replacing n byn + 1 in (18.153) <strong>and</strong> integrating the RHS by parts, we findΓ(n +1)=∫ ∞0x n e −x dx= [ −x n e −x] ∫ ∞∞0 + nx n−1 e −x dx= n∫ ∞00x n−1 e −x dx,from which we obtain the important resultΓ(n +1)=nΓ(n). (18.154)From (18.153), we see that Γ(1) = 1, <strong>and</strong> so, if n is a positive integer,Γ(n +1)=n!. (18.155)In fact, equation (18.155) serves as a definition of the factorial function even <strong>for</strong>non-integer n. For negative n the factorial function is defined by(n + m)!n! =(n + m)(n + m − 1) ···(n +1) , (18.156)where m is any positive integer that makes n + m>0. Different choices of m(> −n) do not lead to different values <strong>for</strong> n!. A plot of the gamma function isgiven in figure 18.9, where it can be seen that the function is infinite <strong>for</strong> negativeinteger values of n, in accordance with (18.156). For an extension of the factorialfunction to complex arguments, see exercise 18.15.By letting x = y 2 in (18.153), we immediately obtain another useful representationof the gamma function given byΓ(n) =2∫ ∞0y 2n−1 e −y2 dy. (18.157)Setting n = 1 2we find the resultΓ ( ∫) ∞ ∫ ∞12 =2 e −y2 dy = e −y2 dy = √ π,0−∞where have used the st<strong>and</strong>ard integral discussed in section 6.4.2. From this result,Γ(n) <strong>for</strong> half-integral n can be found using (18.154). Some immediately derivablefactorial values of half integers are(−32)!=−2√ π,(−12)!=√ π,( 12636) (!=12√ π, 3)2 !=34√ π.


18.12 THE GAMMA FUNCTION AND RELATED FUNCTIONSΓ( n)642−4−3−2−11 23 4n−2−4−6Figure 18.9The gamma function Γ(n).Moreover, it may be shown <strong>for</strong> non-integral n that the gamma function satisfiesthe important identityΓ(n)Γ(1 − n) =πsin nπ . (18.158)This is proved <strong>for</strong> a restricted range of n in the next section, once the betafunction has been introduced.It can also be shown that the gamma function is given byΓ(n +1)= √ (2πn n n e −n 1+ 112n + 1288n 2 − 139 )51 840n 3 + ... = n!,(18.159)which is known as Stirling’s asymptotic series. For large n the first term dominates,<strong>and</strong> son! ≈ √ 2πn n n e −n ; (18.160)this is known as Stirling’s approximation. This approximation is particularly usefulin statistical thermodynamics, when arrangements of a large number of particlesaretobeconsidered.◮Prove Stirling’s approximation n! ≈ √ 2πn n n e −n <strong>for</strong> large n.From (18.153), the extended definition of the factorial function (which is valid <strong>for</strong> n>−1)is given byn! =∫ ∞0x n e −x dx =637∫ ∞0e n ln x−x dx. (18.161)


SPECIAL FUNCTIONSIf we let x = n + y, then(ln x =lnn +ln 1+ y )n=lnn + y n − y22n 2 + y33n 3 − ··· .Substituting this result into (18.161), we obtain∫ ∞[n! = exp n(ln n + y ···) ]−nn − y22n + − n − y dy.2Thus, when n is sufficiently large, we may approximate n! by∫ ∞n! ≈ e n ln n−n e −y2 /(2n) dy = e n ln n−n√ 2πn = √ 2πn n n e −n ,−∞which is Stirling’s approximation (18.160). ◭The beta function is defined by18.12.2 The beta functionB(m, n) =∫ 10x m−1 (1 − x) n−1 dx, (18.162)which converges <strong>for</strong> m>0, n>0, where m <strong>and</strong> n are, in general, real numbers.By letting x =1− y in (18.162) it is easy to show that B(m, n) =B(n, m). Otheruseful representations of the beta function may be obtained by suitable changesof variable. For example, putting x =(1+y) −1 in (18.162), we find that∫ ∞y n−1 dyB(m, n) =. (18.163)0 (1 + y)m+nAlternatively, if we let x =sin 2 θ in (18.162), we obtain immediatelyB(m, n) =2∫ π/20sin 2m−1 θ cos 2n−1 θdθ. (18.164)The beta function may also be written in terms of the gamma function asB(m, n) = Γ(m)Γ(n)Γ(m + n) . (18.165)◮Prove the result (18.165).Using (18.157), we haveΓ(n)Γ(m) =4=4∫ ∞0∫ ∞ ∫ ∞0 0x 2n−1 e −x2 dx∫ ∞0y 2m−1 e −y2 dyx 2n−1 y 2m−1 e −(x2 +y 2) dx dy.638


18.12 THE GAMMA FUNCTION AND RELATED FUNCTIONSChanging variables to plane polar coordinates (ρ, φ) givenbyx = ρ cos φ, y = ρ sin φ, weobtainΓ(n)Γ(m) =4=4∫ π/2 ∫ ∞0 0∫ π/20ρ 2(m+n−1) e −ρ2 sin 2m−1 φ cos 2n−1 φ ρ dρ dφsin 2m−1 φ cos 2n−1 φdφ∫ ∞0ρ 2(m+n)−1 e −ρ2 dρ= B(m, n)Γ(m + n),where in the last line we have used the results (18.157) <strong>and</strong> (18.164). ◭The result (18.165) is useful in proving the identity (18.158) satisfied by thegamma function, since∫ ∞y n−1 dyΓ(n)Γ(1 − n) =B(1 − n, n) =0 1+y ,where, in the second equality, we have used the integral representation (18.163).For 0


SPECIAL FUNCTIONSwhich is the required result. ◭We note that is it conventional to define, in addition, the functionsγ(a, x)Γ(a, x)P (a, x) ≡ , Q(a, x) ≡Γ(a) Γ(a) ,which are also often called incomplete gamma functions; it is clear that Q(a, x) =1 − P (a, x).18.12.4 The error functionFinally, we mention the error function, which is encountered in probability theory<strong>and</strong> in the solutions of some partial differential equations. The error function isrelated to the incomplete gamma function by erf(x) =γ( 1 2 ,x2 )/ √ π <strong>and</strong> is thusgiven byerf(x) = √ 2 ∫ xe −u2 du =1− 2 ∫ ∞√ e −u2 du. (18.167)π πFrom this definition we can easily see that0erf(0) = 0, erf(∞) =1, erf(−x) =−erf(x).By making the substitution y = √ 2u in (18.167), we find√ ∫ √22xerf(x) = e −y2 /2 dy.π 0The cumulative probability function Φ(x) <strong>for</strong> the st<strong>and</strong>ard Gaussian distribution(discussed in section 30.9.1) may be written in terms of the error function asfollows:Φ(x) = √ 1 ∫ xe −y2 /2 dy2π −∞= 1 2 + √ 1 ∫ xe −y2 /2 dy2π 0= 1 2 + 1 ( ) x√22 erf .It is also sometimes useful to define the complementary error functionerfc(x) =1− erf(x) = √ 2 ∫ ∞e −u2 du = Γ( 1 2 ,x2 )√ . (18.168)π πxx18.1 Use the explicit expressions18.13 Exercises640


18.13 EXERCISES√√Y0 0 = 1, Y 04π 1 = 3cos θ,4π√√Y ±1301= ∓ sin θ exp(±iφ), Y8π 2 = 5(3 16π cos2 θ − 1),√√Y ±1±22= ∓ sin θ cos θ exp(±iφ), Y2=158πto verify <strong>for</strong> l =0, 1, 2that1532π sin2 θ exp(±2iφ),l∑|Yl m (θ, φ)| 2 2l +1=4π ,m=−l<strong>and</strong> so is independent of the values of θ <strong>and</strong> φ. Thisistrue<strong>for</strong>anyl, buta general proof is more involved. This result helps to reconcile intuition withthe apparently arbitrary choice of polar axis in a general quantum mechanicalsystem.18.2 Express the functionf(θ, φ) =sinθ[sin 2 (θ/2) cos φ + i cos 2 (θ/2) sin φ]+sin 2 (θ/2)as a sum of spherical harmonics.18.3 Use the generating function <strong>for</strong> the Legendre polynomials P n (x) to show that∫ 1P 2n+1 (x) dx =(−1) n (2n)!02 2n+1 n!(n +1)!<strong>and</strong> that, except <strong>for</strong> the case n =0,∫ 10P 2n (x) dx =0.18.4 Carry through the following procedure as a proof of the result∫ 1I n = P n (z)P n (z) dz = 2−12n +1 .(a) Square both sides of the generating-function definition of the Legendrepolynomials,∞∑(1 − 2zh + h 2 ) −1/2 = P n (z)h n .(b) Express the RHS as a sum of powers of h, obtaining expressions <strong>for</strong> thecoefficients.(c) Integrate the RHS from −1 to 1 <strong>and</strong> use the orthogonality property of theLegendre polynomials.(d) Similarly integrate the LHS <strong>and</strong> exp<strong>and</strong> the result in powers of h.(e) Compare coefficients.18.5 The Hermite polynomials H n (x) may be defined byShow thatΦ(x, h) =exp(2xh − h 2 )=n=0∞∑n=0∂ 2 Φ ∂Φ ∂Φ− 2x +2h∂x2 ∂x ∂h =0,6411n! H n(x)h n .


SPECIAL FUNCTIONS<strong>and</strong> hence that the H n (x) satisfy the Hermite equationy ′′ − 2xy ′ +2ny =0,where n is an integer ≥ 0.Use Φ to prove that(a) H n(x) ′ =2nH n−1 (x),(b) H n+1 (x) − 2xH n (x)+2nH n−1 (x) =0.18.6 A charge +2q is situated at the origin <strong>and</strong> charges of −q are situated at distances±a from it along the polar axis. By relating it to the generating function <strong>for</strong> theLegendre polynomials, show that the electrostatic potential Φ at a point (r, θ, φ)with r>ais given byΦ(r, θ, φ) =2q ∞∑ ( a) 2sP2s (cos θ).4πɛ 0 r rs=118.7 For the associated Laguerre polynomials, carry through the following exercises.(a) Prove the Rodrigues’ <strong>for</strong>mulaL m n (x) = ex x −m d nn! dx n (xn+m e −x ),taking the polynomials to be defined byn∑L m n (x) = (−1) k (n + m)!k!(n − k)!(k + m)! xk .k=0(b) Prove the recurrence relations(n +1)L m n+1(x) =(2n + m +1− x)L m n (x) − (n + m)L m n−1(x),x(L m n ) ′ (x) =nL m n (x) − (n + m)L m n−1(x),but this time taking the polynomial as defined byL m n (x) =(−1) m dmdx L n+m(x)mor the generating function.18.8 The quantum mechanical wavefunction <strong>for</strong> a one-dimensional simple harmonicoscillator in its nth energy level is of the <strong>for</strong>mψ(x) =exp(−x 2 /2)H n (x),where H n (x) isthenth Hermite polynomial. The generating function <strong>for</strong> thepolynomials is∞∑ HG(x, h) =e 2hx−h2 n (x)= h n .n!n=0(a) Find H i (x) <strong>for</strong>i =1, 2, 3, 4.(b) Evaluate by direct calculation∫ ∞−∞e −x2 H p (x)H q (x) dx,(i) <strong>for</strong> p =2,q = 3; (ii) <strong>for</strong> p =2,q = 4; (iii) <strong>for</strong> p = q = 3. Check youranswers against the expected values 2 p p! √ πδ pq .642


18.13 EXERCISES[ You will find it convenient to use∫ ∞x 2n e −x2 dx = (2n)!√ π−∞2 2n n!<strong>for</strong> integer n ≥ 0. ]18.9 By initially writing y(x) asx 1/2 f(x) <strong>and</strong> then making subsequent changes ofvariable, reduce Stokes’ equation,d 2 y+ λxy =0,dx2 to Bessel’s equation. √Hence show that a solution that is finite at x =0isamultiple of x 1/2 J 1/3 ( 2 λx3 ).318.10 By choosing a suitable <strong>for</strong>m <strong>for</strong> h in their generating function,[ ( zG(z,h) =exp h − 1 )] ∞∑= J n (z)h n ,2 hn=−∞show that integral repesentations of the Bessel functions of the first kind aregiven, <strong>for</strong> integral m, by∫ 2πJ 2m (z) = (−1)m cos(z cos θ) cos2mθ dθ m ≥ 1,π 0∫ 2πJ 2m+1 (z) = (−1)m+1 cos(z cos θ) sin(2m +1)θdθ m≥ 0.π 018.11 Identify the series <strong>for</strong> the following hypergeometric functions, writing them interms of better known functions:(a) F(a, b, b; z),(b) F(1, 1, 2; −x),(c) F( 1 , 1, 3 ; 2 2 −x2 ),(d) F( 1 , 1 , 3 ; 2 2 2 x2 ),(e) F(−a, a, 1 2 ;sin2 x); this is a much more difficult exercise.18.12 By making the substitution z =(1− x)/2 <strong>and</strong> suitable choices <strong>for</strong> a, b <strong>and</strong> c,convert the hypergeometric equation,z(1 − z) d2 udu+[c − (a + b +1)z ] − abu =0,dz2 dzinto the Legendre equation,(1 − x 2 ) d2 y dy− 2x + l(l +1)y =0.dx2 dxHence, using the hypergeometric series, generate the Legendre polynomials P l (x)<strong>for</strong> the integer values l =0, 1, 2, 3. Comment on their normalisations.18.13 Find a change of variable that will allow the integral∫ ∞√u − 1I =1 (u +1) du 2to be expressed in terms of the beta function, <strong>and</strong> so evaluate it.18.14 Prove that, if m <strong>and</strong> n are both greater than −1, then∫ ∞u mI =0 (au 2 + b) du = Γ[ 1 (m 2 +1)]Γ[1 (n +1)]2 (m+n+2)/2 2a (m+1)/2 b (n+1)/2 Γ[ 1 (m + n +2)].2643


SPECIAL FUNCTIONSDeduce the value of∫ ∞(u +2) 2J =du.0 (u 2 +4)5/218.15 The complex function z! is defined byz! =∫ ∞0u z e −u du<strong>for</strong> Re z>−1.For Re z ≤−1 it is defined by(z + n)!z! =(z + n)(z + n − 1) ···(z +1) ,where n is any (positive) integer > −Re z. Being the ratio of two polynomials, z!is analytic everywhere in the finite complex plane except at the poles that occurwhen z is a negative integer.(a) Show that the definition of z! <strong>for</strong>Rez ≤−1 is independent of the value ofn chosen.(b) Prove that the residue of z! at the pole z = −m, wherem is an integer > 0,is (−1) m−1 /(m − 1)!.18.16 For −1 < Re z0, occurs in some statistical mechanics problems. By first consideringthe integral∫ ∞J = e iu(k+ia) du,<strong>and</strong> a suitable variation of it, show that I =(π/a) exp(a 2 )erfc(a), where erfc(x)is the complementary error function.18.18 Consider two series expansions of the error function as follows.(a) Obtain a series expansion of the error function erf(x) in ascending powersof x. How many terms are needed to give a value correct to four significantfigures <strong>for</strong> erf(1)?(b) Obtain an asymptotic expansion that can be used to estimate erfc(x) <strong>for</strong>large x (> 0) in the <strong>for</strong>m of a seriesaerfc(x) =R(x) =e −x2 nx . n n=0Consider what bounds can be put on the estimate <strong>and</strong> at what point theinfinite series should be terminated in a practical estimate. In particular,estimate erfc(1) <strong>and</strong> test the answer <strong>for</strong> compatibility with that in part (a).18.19 For the functions M(a, c; z) that are the solutions of the confluent hypergeometricequation,0644∞∑


18.13 EXERCISES(a) use their series representation to prove thatb d M(a, c; z) =aM(a +1,c+1;z);dz(b) use an integral representation to prove thatM(a, c; z) =e z M(c − a, c; −z).18.20 The Bessel function J ν (z) can be considered as a special case of the solutionM(a, c; z) of the confluent hypergeometric equation, the connection beingM(a, ν +1;−z/a)lim= z −ν/2 J ν (2 √ z).a→∞ Γ(ν +1)Prove this equality by writing each side in terms of an infinite series <strong>and</strong> showingthat the series are the same.18.21 Find the differential equation satisfied by the function y(x) defined by∫ xy(x) =Ax −n e −t t n−1 dt ≡ Ax −n γ(n, x),0<strong>and</strong>, by comparing it with the confluent hypergeometric function, express y asa multiple of the solution M(a, c; z) of that equation. Determine the value of Athat makes y equal to M.18.22 Show, from its definition, that the Bessel function of the second kind, <strong>and</strong> ofintegral order ν, can be written asY ν (z) = 1 [ ∂Jµ (z)− (−1) ν ∂J ]−µ(z).π ∂µ∂µµ=νUsing the explicit series expression <strong>for</strong> J µ (z), show that ∂J µ (z)/∂µ can be writtenas( z)J ν (z)ln + g(ν,z),2<strong>and</strong> deduce that Y ν (z) can be expressed asY ν (z) = 2 ( z)π J ν(z)ln + h(ν,z),2where h(ν,z), like g(ν,z), is a power series in z.18.23 Prove two of the properties of the incomplete gamma function P (a, x 2 ) as follows.(a) By considering its <strong>for</strong>m <strong>for</strong> a suitable value of a, show that the error functioncan be expressed as a particular case of the incomplete gamma function.(b) The Fresnel integrals, of importance in the study of the diffraction of light,are given by∫ x ( π)∫ x ( π)C(x) = cos2 t2 dt, S(x) = sin2 t2 dt.0Show that they can be expressed in terms of the error function by[ √ ] πC(x)+iS(x) =A erf2 (1 − i)x ,where A is a (complex) constant, which you should determine. Hence expressC(x)+iS(x) in terms of the incomplete gamma function.0645


SPECIAL FUNCTIONS18.24 The solutions y(x, a) of the equationd 2 ydx − ( 1 2 4 x2 + a)y =0 (∗)are known as parabolic cylinder functions.(a) If y(x, a) is a solution of (∗), determine which of the following are alsosolutions: (i) y(a, −x), (ii) y(−a, x), (iii) y(a, ix) <strong>and</strong>(iv)y(−a, ix).(b) Show that one solution of (∗), even in x, isy 1 (x, a) =e −x2 /4 M( 1 a + 1 , 1 , 1 2 4 2 2 x2 ),where M(α, c, z) is the confluent hypergeometric function satisfyingz d2 M+(c − z) dM − αM =0.dz 2 dzYou may assume (or prove) that a second solution, odd in x, isgivenbyy 2 (x, a) =xe −x2 /4 M( 1 a + 3 , 3 , 1 2 4 2 2 x2 ).(c) Find, as an infinite series, an explicit expression <strong>for</strong> e x2 /4 y 1 (x, a).(d) Using the results from part (a), show that y 1 (x, a) can also be written asy 1 (x, a) =e x2 /4 M(− 1 a + 1 , 1 , − 1 2 4 2 2 x2 ).(e) By making a suitable choice <strong>for</strong> a deduce that()∞∑ b n x 2n∞∑1+(2n)! = ex2 /2 (−1) n b n x 2n1+,(2n)!n=1n=1where b n = ∏ nr=1 (2r − 3 ). 218.14 Hints <strong>and</strong> answers18.1 Note that taking the square of the modulus eliminates all mention of φ.18.3 Integrate both sides of the generating function definition from x =0tox =1,<strong>and</strong> then exp<strong>and</strong> the resulting term, (1+h 2 ) 1/2 , using a binomial expansion. Showthat 1/2 C m can be written as [ (−1) m−1 (2m − 2)! ]/[2 2m−1 m!(m − 1)! ].18.5 Prove the stated equation using the explicit closed <strong>for</strong>m of the generating function.Then substitute the series <strong>and</strong> require the coefficient of each power of h to vanish.(b) Differentiate result (a) <strong>and</strong> then use (a) again to replace the derivatives.18.7 (a) Write the result of using Leibnitz’ theorem on the product of x n+m <strong>and</strong> e −x as afinite sum, evaluate the separated derivatives, <strong>and</strong> then re-index the summation.(b) For the first recurrence relation, differentiate the generating function withrespect to h <strong>and</strong> then use the generating function again to replace the exponential.Equating coefficients of h n then yields the result. For the second, differentiate thecorresponding relationship <strong>for</strong> the ordinary Laguerre polynomials m times.18.9 x 2 f ′′ + xf ′ +(λx 3 − 1 4 )f = 0. Then, in turn, set x3/2 = u, <strong>and</strong> 2 3 λ1/2 u = v; thenvsatisfies Bessel’s equation with ν = 1 3 .18.11 (a) (1 − z) −a .(b)x −1 ln(1 + x). (c) Compare the calculated coefficients with thoseof tan −1 x. F( 1 2 , 1, 3 2 ; −x2 )=x −1 tan −1 x. (d) x −1 sin −1 x. (e) Note that a termcontaining x 2n can only arise from the first n + 1 terms of an expansion in powersof sin 2 x; make a few trials. F(−a, a, 1 2 ;sin2 x)=cos2ax.18.13 Looking <strong>for</strong> f(x) =u such that u + 1 is an inverse power of x with f(0) = ∞ <strong>and</strong>f(1) = 1 leads to f(x) =2x −1 − 1. I = B( 1 2 , 3 2 )/√ 2=π/(2 √ 2).646


18.14 HINTS AND ANSWERS18.15 (a) Show that the ratio of two definitions based on m <strong>and</strong> n, withm>n>−Re z,is unity, independent of the actual values of m <strong>and</strong> n.(b) Consider the limit as z →−m of (z + m)z!, with the definition of z! basedonn where n>m.18.17 Express the integr<strong>and</strong> in partial fractions <strong>and</strong> use J, as given, <strong>and</strong> J ′ =∫ ∞0exp[ −iu(k −ia)]du to express I as the sum of two double integral expressions.Reduce them using the st<strong>and</strong>ard Gaussian integral, <strong>and</strong> then make a change ofvariable 2v = u +2a.18.19 (b) Using the representationΓ(b)M(a, b; z) =e zt t a−1 (1 − t) b−a−1 dtΓ(b − a)Γ(a) 0allows the equality to be established, without actual integration, by changing theintegration variable to s =1− t.18.21 Calculate y ′ (x) <strong>and</strong>y ′′ (x) <strong>and</strong> then eliminate x −1 e −x to obtain xy ′′ +(n+1+x)y ′ +ny =0;M(n, n +1;−x). Comparing the expansion of the hypergeometric serieswith the result of term by term integration of the expansion of the integr<strong>and</strong>shows that A = n.18.23 (a) If the dummy variable in the incomplete gamma function is t, make the changeof variable y =+ √ t.Nowchoosea so that 2(a − 1) + 1 = 0; erf(x) =P ( 1 2 ,x2 ).(b) Change the integration √ variable u in the st<strong>and</strong>ard representation of the RHSto s, givenbyu = 1 2 π(1 − i)s, <strong>and</strong> note that (1 − i) 2 = −2i. A =(1+i)/2. Frompart (a), C(x)+iS(x) = 1 (1 + i)P ( 1 , − 1 πi 2 2 2 x2 ).∫ 1647


19Quantum operatorsAlthough the previous chapter was principally concerned with the use of linearoperators <strong>and</strong> their eigenfunctions in connection with the solution of givendifferential equations, it is of interest to study the properties of the operatorsthemselves <strong>and</strong> determine which of them follow purely from the nature of theoperators, without reference to specific <strong>for</strong>ms of eigenfunctions.19.1 Operator <strong>for</strong>malismThe results we will obtain in this chapter have most of their applications in thefield of quantum mechanics <strong>and</strong> our descriptions of the methods will reflect this.In particular, when we discuss a function ψ that depends upon variables such asspace coordinates <strong>and</strong> time, <strong>and</strong> possibly also on some non-classical variables, ψwill usually be a quantum-mechanical wavefunction that is being used to describethe state of a physical system. For example, the value of |ψ| 2 <strong>for</strong> a particularset of values of the variables is interpreted in quantum mechanics as being theprobability that the system’s variables have that set of values.To this end, we will be no more specific about the functions involved thanattaching just enough labels to them that a particular function, or a particularset of functions, is identified. A convenient notation <strong>for</strong> this kind of approachis that already hinted at, but not specifically stated, in subsection 17.1, wherethe definition of an inner product is given. This notation, often called the Diracnotation, denotes a state whose wavefunction is ψ by | ψ〉; sinceψ belongs to avector space of functions, | ψ〉 is known as a ket vector. Ket vectors, or simply kets,must not be thought of as completely analogous to physical vectors. Quantummechanics associates the same physical state with ke iθ | ψ〉 as it does with | ψ〉<strong>for</strong> all real k <strong>and</strong> θ <strong>and</strong> so there is no loss of generality in taking k as 1 <strong>and</strong> θas 0. On the other h<strong>and</strong>, the combination c 1 | ψ 1 〉 + c 2 | ψ 2 〉,where| ψ 1 〉 <strong>and</strong> | ψ 2 〉648


19.1 OPERATOR FORMALISMrepresent different states, is a ket that represents a continuum of different statesas the complex numbers c 1 <strong>and</strong> c 2 are varied.If we need to specify a state more closely – say we know that it correspondsto a plane wave with a wave number whose magnitude is k – then we indicatethis with a label; the corresponding ket vector would be written as | k〉. If we alsoknew the direction of the wave then | k〉 would be the appropriate <strong>for</strong>m. Clearly,in general, the more labels we include, the more precisely the corresponding stateis specified.The Dirac notation <strong>for</strong> the Hermitian conjugate (dual vector) of the ket vector| ψ〉 is written as 〈ψ| <strong>and</strong> is known as a bra vector; the wavefunction describing thisstate is ψ ∗ , the complex conjugate of ψ. The inner product of two wavefunctions∫ψ ∗ φdvis then denoted by 〈ψ| φ〉 or, more generally if a non-unit weight functionρ is involved, by〈ψ| ρ| φ〉,evaluated as∫ψ ∗ (r)φ(r)ρ(r) dr. (19.1)Given the (contrived) names <strong>for</strong> the two sorts of vectors, an inner product like〈ψ| φ〉 becomes a particular type of ‘bra(c)ket’. Despite its somewhat whimsicalconstruction, this type of quantity has a fundamental role to play in the interpretationof quantum theory, because expectation values, probabilities <strong>and</strong> transitionrates are all expressed in terms of them. For physical states the inner product ofthe corresponding ket with itself, with or without an explicit weight function, isnon-zero, <strong>and</strong> it is usual to take〈ψ|ψ〉 =1.Although multiplying a ket vector by a constant does not change the statedescribed by the vector, acting upon it with a more general linear operator Aresults (in general) in a ket describing a different state. For example, if ψ is a statethat is described in one-dimensional x-space by the wavefunction ψ(x) =exp(−x 2 )<strong>and</strong> A is the differential operator ∂/∂x, then| ψ 1 〉 = A| ψ〉 ≡|Aψ〉is the ket associated with the state whose wavefunction is ψ 1 (x) =−2x exp(−x 2 ),clearly a different state. This allows us to attach a meaning to an expression suchas 〈φ|A| ψ〉 through the equation〈φ|A| ψ〉 = 〈φ| ψ 1 〉, (19.2)i.e. it is the inner product of | ψ 1 〉 <strong>and</strong> | φ〉. We have already used this notation inequation (19.1), but there the effect of the operator A was merely multiplicationby a weight function.If it should happen that the effect of an operator acting upon a particular ket649


QUANTUM OPERATORSis to produce a scalar multiple of that ket, i.e.A| ψ〉 = λ| ψ〉, (19.3)then, just as <strong>for</strong> matrices <strong>and</strong> differential equations, | ψ〉 is called an eigenket or,more usually, an eigenstate of A, with corresponding eigenvalue λ; tomarkthisspecial property the state will normally be denoted by | λ〉, rather than by themore general | ψ〉. Taking the Hermitian conjugate of this ket vector eigenequationgives a bra vector equation,〈ψ|A † = λ ∗ 〈ψ|. (19.4)It should be noted that the complex conjugate of the eigenvalue appears in thisequation. Should the action of A on |ψ〉 produce an unphysical state (usuallyone whose wavefunction is identically zero, <strong>and</strong> is there<strong>for</strong>e unacceptable as aquantum-mechanical wavefunction because of the required probability interpretation)we denote the result either by 0 or by the ket vector |∅〉 according tocontext. Formally, |∅〉 can be considered as an eigenket of any operator, but one<strong>for</strong> which the eigenvalue is always zero.If an operator A is Hermitian (A † = A) then its eigenvalues are real <strong>and</strong>the eigenstates can be chosen to be orthogonal; this can be shown in the sameway as in chapter 17 (but using a different notation). As indicated there, thereality of their eigenvalues is one reason why Hermitian operators <strong>for</strong>m thebasis of measurement in quantum mechanics; in that <strong>for</strong>mulation of physics, theeigenvalues of an operator are the only possible values that can be obtained whena measurement of the physical quantity corresponding to the operator is made.Actual individual measurements must always result in real values, even if theyare combined in a complex <strong>for</strong>m (x + iy or re iθ ) <strong>for</strong> final presentation or analysis,<strong>and</strong> using only Hermitian operators ensures this. The proof of the reality of theeigenvalues using the Dirac notation is given below in a worked example.In the same notation the Hermitian property of an operator A is representedby the double equality〈Aφ|ψ〉 = 〈φ|A| ψ〉 = 〈φ|Aψ〉.It should be remembered that the definition of an Hermitian operator involvesspecifying boundary conditions that the wavefunctions considered must satisfy.Typically, they are that the wavefunctions vanish <strong>for</strong> large values of the spatialvariables upon which they depend; this deals with most physical systems sincethey are nearly all <strong>for</strong>mally infinite in extent. Some model systems require thewavefunction to be periodic or to vanish at finite values of a spatial variable.Depending on the nature of the physical system, the eigenvalues of a particularlinear operator may be discrete, part of a continuum, or a mixture of both. Forexample, the energy levels of the bound proton–electron system (the hydrogenatom) are discrete, but if the atom is ionised <strong>and</strong> the electron is free, the energy650


19.1 OPERATOR FORMALISMspectrum of the system is continuous. This system has discrete negative <strong>and</strong>continuous positive eigenvalues <strong>for</strong> the operator corresponding to the total energy(the Hamiltonian).◮Using the Dirac notation, show that the eigenvalues of an Hermitian operator are real.Let | a〉 be an eigenstate of Hermitian operator A corresponding to eigenvalue a, thenHence,⇒⇒A| a〉 = a| a〉,〈a|A| a〉 = 〈a|a| a〉 = a〈a| a〉,<strong>and</strong>〈a|A † = a ∗ 〈a|,〈a|A † | a〉 = a ∗ 〈a| a〉,〈a|A| a〉 = a ∗ 〈a| a〉, since A is Hermitian.(a − a ∗ )〈a| a〉 = 0,⇒ a = a ∗ , since 〈a| a〉 ≠0.Thus a is real. ◭It is not our intention to describe the complete axiomatic basis of quantummechanics, but rather to show what can be learned about linear operators, <strong>and</strong>in particular about their eigenvalues, without recourse to explicit wavefunctionson which the operators act.Be<strong>for</strong>e we proceed to do that, we close this subsection with a number of results,expressed in Dirac notation, that the reader should verify by inspection or byfollowing the lines of argument sketched in the statements. Where a sum overa complete set of eigenvalues is shown, it should be replaced by an integral <strong>for</strong>those parts of the eigenvalue spectrum that are continuous. With the notationthat |a n 〉 is an eigenstate of Hermitian operator A with non-degenerate eigenvaluea n (or, if a n is k-fold degenerate, then a set of k mutually orthogonal eigenstateshas been constructed <strong>and</strong> the states relabelled), we have the following results.A| a n 〉 = a n | a n 〉,〈a m |a n 〉 = δ mn (orthonormality of eigenstates), (19.5)A(c n | a n 〉 + c m | a m 〉)=c n a n | a n 〉 + c m a m | a m 〉 (linearity). (19.6)The definitions of the sum <strong>and</strong> product of two operators are(A + B)| ψ〉 ≡A| ψ〉 + B| ψ〉, (19.7)AB| ψ〉 ≡A(B| ψ〉) (≠ BA| ψ〉 in general), (19.8)⇒ A p | a n 〉 = a p n| a n 〉. (19.9)651


QUANTUM OPERATORSIf A| a n 〉 = a| a n 〉 <strong>for</strong> all N 1 ≤ n ≤ N 2 ,then∑N 2| ψ〉 = d n | a n 〉 satisfies A| ψ〉 = a| ψ〉 <strong>for</strong> any set of d i .n=N 1For a general state | ψ〉,| ψ〉 =∞∑c n | a n 〉,wherec n = 〈a n |ψ〉. (19.10)n=0This can also be expressed as the operator identity,∞∑1= | a n 〉〈a n |, (19.11)in the sense thatn=0n=0∞∑∞∑| ψ〉 =1| ψ〉 = | a n 〉〈a n |ψ〉 = c n | a n 〉.It also follows that( ∞)(∑∞)∑∞∑∞∑1=〈ψ|ψ〉 = c ∗ m〈a m | c n |a n 〉 = c ∗ mc n δ mn = |c n | 2 .m=0n=0m,nn=0 (19.12)Similarly, the expectation value of the physical variable corresponding to A is∞∑∞∑〈ψ|A| ψ〉 = c ∗ m〈a m | A| a n 〉c n = c ∗ m〈a m | a n | a n 〉c nm,nm,nm,nn=0∞∑∞∑= c ∗ mc n a n δ mn = |c n | 2 a n . (19.13)n=019.1.1 Commutation <strong>and</strong> commutatorsAs has been noted above, the product AB of two linear operators may or maynot be equal to the product BA. ThatisAB| ψ〉 is not necessarily equal to BA| ψ〉.If A <strong>and</strong> B are both purely multiplicative operators, multiplication by f(r)<strong>and</strong> g(r) say, then clearly the order of the operations is immaterial, the result| f(r)g(r)ψ〉 being obtained in both cases. However, consider a case in which Ais the differential operator ∂/∂x <strong>and</strong> B is the operator ‘multiply by x’. Then thewavefunction describing AB| ψ〉 is∂(xψ(x)) = ψ(x)+x∂ψ∂x ∂x ,652


19.1 OPERATOR FORMALISMwhilstthat<strong>for</strong>BA| ψ〉 is simplywhich is not the same.If the resultx ∂ψ∂x ,AB| ψ〉 = BA| ψ〉is true <strong>for</strong> all ket vectors | ψ〉, thenA <strong>and</strong> B are said to commute; otherwise theyare non-commuting operators.A convenient way to express the commutation properties of two linear operatorsis to define their commutator, [ A, B ], by[ A, B ] | ψ〉 ≡AB| ψ〉−BA| ψ〉. (19.14)Clearly two operators that commute have a zero commutator. But, <strong>for</strong> the examplegiven above we have that[ ∂∂x ,x ]ψ(x) =(ψ(x)+x ∂ψ ) (− x ∂ψ )= ψ(x) =1× ψ∂x ∂xor, more simply, that[ ∂∂x ,x ]= 1; (19.15)in words, the commutator of the differential operator ∂/∂x <strong>and</strong> the multiplicativeoperator x is the multiplicative operator 1. It should be noted that the order ofthe linear operators is important <strong>and</strong> that[ A, B ] = − [ B,A] . (19.16)Clearly any linear operator commutes with itself <strong>and</strong> some other obvious zerocommutators (when operating on wavefunctions with ‘reasonable’ properties) are:[ A, I ] , where I is the identity operator;[ A n ,A m ] , <strong>for</strong> any positive integers n <strong>and</strong> m;[ A, p(A) ] , where p(x) is any polynomial in x;[ A, c ] , where A is any linear operator <strong>and</strong> c is any constant;[ f(x),g(x) ] , where the functions are mutiplicative;[ A(x),B(y) ] , where the operators act on different variables, with[ ] ∂∂x , ∂as a specific example.∂y653


QUANTUM OPERATORSSimple identities amongst commutators include the following:[ A, B + C ] = [ A, B ] + [ A, C ] , (19.17)[ A + B,C ] = [ A, C ] + [ B,C ] , (19.18)[ A, BC ] = ABC − BCA + BAC − BAC=(AB − BA)C + B(AC − CA)= [ A, B ] C + B [ A, C ] , (19.19)[ AB, C ] = A [ B,C ] + [ A, C ] B. (19.20)◮If A <strong>and</strong> B are two linear operators that both commute with their commutator, prove that[ A, B n ] = nB n−1 [ A, B ] <strong>and</strong> that [ A n ,B] = nA n−1 [ A, B ].Define C n by C n = [ A, B n ]. We aim to find a reduction <strong>for</strong>mula <strong>for</strong> C n :C n = [ A, B B ] n−1= [ A, B ] B n−1 + B [ A, B ] n−1 , using (19.19),= B n−1 [ A, B ] + B [ A, B ] n−1 ,since[[A, B ] ,B] =0,= B n−1 [ A, B ] + BC n−1 , the required reduction <strong>for</strong>mula,= B n−1 [ A, B ] + B{B n−2 [ A, B ] + BC n−2 }, applying the <strong>for</strong>mula,=2B n−1 [ A, B ] + B 2 C n−2= ···= nB n−1 [ A, B ] + B n C 0 .However, C 0 = [ A, I ] =0<strong>and</strong>soC n = nB n−1 [ A, B ].Using equation (19.16) <strong>and</strong> interchanging A <strong>and</strong> B in the result just obtained, we findas stated in the question. ◭[ A n ,B] = − [ B,A n ] = −nA n−1 [ B,A] = nA n−1 [ A, B ] ,As the power of a linear operator can be defined, so can its exponential;this situation parallels that <strong>for</strong> matrices, which are of course a particular set ofoperators that act upon state functions represented by vectors. The definitionfollows that <strong>for</strong> the exponential of a scalar or matrix, namelyexp A =∞∑n=0A nn! . (19.21)Related functions of A, suchassinA <strong>and</strong> cos A, can be defined in a similar way.Since any linear operator commutes with itself, when two functions of itare combined in some way, the result takes a <strong>for</strong>m similar to that <strong>for</strong> thecorresponding functions of scalar quantities. Consider, <strong>for</strong> example, the functionf(A) defined by f(A) =2sinA cos A. Expressing sin A <strong>and</strong> cos A in terms of their654


19.1 OPERATOR FORMALISMdefining series, we havef(A) =2∞∑m=0(−1) m A 2m+1(2m +1)!∞∑n=0(−1) n A 2n.(2n)!Writing m + n as r <strong>and</strong> replacing n by s, we havewherec r =f(A) =2∞∑( r∑A 2r+1r=0 s=0∞∑=2 (−1) r c r A 2r+1 ,r∑s=0r=0(−1) r−s(2r − 2s +1)!1(2r − 2s + 1)! (2s)! = 1(2r +1)!)(−1) s(2s)!r∑2r+1 C 2s .By adding the binomial expansions of 2 2r+1 =(1+1) 2r+1 <strong>and</strong> 0 = (1 − 1) 2r+1 ,itcan easily be shown thatIt then follows that2 2r+1 =2s=0s=0r∑2r+1 2 2rC 2s ⇒ c r =(2r +1)! .2sinA cos A =2∞∑r=0(−1) r A 2r+1 2 2r(2r +1)!=∞∑r=0(−1) r (2A) 2r+1(2r +1)!= sin 2A,a not unexpected result.However, if two (or more) linear operators that do not commute are involved,combining functions of them is more complicated <strong>and</strong> the results less intuitivelyobvious. We take as a particular case the product of two exponential functions<strong>and</strong>, even then, take the simplified case in which each linear operator commuteswith their commutator (so that we may use the results from the previous workedexample).◮If A <strong>and</strong> B are two linear operators that both commute with their commutator, show thatexp(A) exp(B) =exp(A + B + 1 [ A, B ] ).2We first find the commutator of A <strong>and</strong> exp λB, whereλ is a scalar quantity introduced <strong>for</strong>655


QUANTUM OPERATORSlater algebraic convenience:[ ] [∞∑A, eλB= A,==∞∑n=0∞∑n=1n=0(λB) nn!]∞∑=n=0λ nn! [ A, Bn ]λ nn! nBn−1 [ A, B ] , using the earlier result,λ nn! nBn−1 [ A, B ]∞∑ λ m B m= λ [ A, B ] ,writingm = n − 1,m!m=0= λe λB [ A, B ] .Now consider the derivative with respect to λ of the functionf(λ) =e λA e λB e −λ(A+B) .In the following calculation we use the fact that the derivative of e λC is Ce λC ; this is thesame as e λC C, since any two functions of the same operator commute. Differentiating thethree-factor product givesdfdλ = eλA Ae λB e −λ(A+B) + e λA e λB Be −λ(A+B) + e λA e λB (−A − B)e −λ(A+B)= e λA (e λB A + λe λB [ A, B ] )e −λ(A+B) + e λA e λB Be −λ(A+B)− e λA e λB Ae −λ(A+B) − e λA e λB Be −λ(A+B)= e λA λe λB [ A, B ] e −λ(A+B)= λ [ A, B ] f(λ).In the second line we have used the result obtained above to replace Ae λB ,<strong>and</strong>inthelastline have used the fact that [ A, B ] commutes with each of A <strong>and</strong> B, <strong>and</strong> hence with anyfunction of them.Integrating this scalar differential equation with respect to λ <strong>and</strong> noting that f(0) = 1,we obtainln f = 1 2 λ2 [ A, B ] ⇒ e λA e λB e −λ(A+B) = f(λ) =e 1 2 λ2 [ A,B ] .Finally, post-multiplying both sides of the equation by e λ(A+B) <strong>and</strong> setting λ =1yieldse A e B = e 1 2 [ A,B ]+A+B . ◭19.2 Physical examples of operatorsWe now turn to considering some of the specific linear operators that play apart in the description of physical systems. In particular, we will examine theproperties of some of those that appear in the quantum-mechanical descriptionof the physical world.As stated earlier, the operators corresponding to physical observables are restrictedto Hermitian operators (which have real eigenvalues) as this ensures thereality of predicted values <strong>for</strong> experimentally measured quantities. The two basic656


19.2 PHYSICAL EXAMPLES OF OPERATORSquantum-mechanical operators are those corresponding to position r <strong>and</strong> momentump. One prescription <strong>for</strong> making the transition from classical to quantummechanics is to express classical quantities in terms of these two variables inCartesian coordinates <strong>and</strong> then make the component by component substitutionsr → multiplicative operator r <strong>and</strong> p → differential operator − i∇.(19.22)This generates the quantum operators corresponding to the classical quantities.For the sake of completeness, we should add that if the classical quantity containsa product of factors whose corresponding operators A <strong>and</strong> B do not commute,then the operator 1 2(AB + BA) is to be substituted <strong>for</strong> the product.The substitutions (19.22) invoke operators that are closely connected with thetwo that we considered at the start of the previous subsection, namely x <strong>and</strong>∂/∂x. One,x, corresponds exactly to the x-component of the prescribed quantumposition operator; the other, however, has been multiplied by the imaginaryconstant −i, where is the Planck constant divided by 2π. This has the (subtle)effect of converting the differential operator into the x-component of an Hermitianoperator; this is easily verified using integration by parts to show that it satisfiesequation (17.16). Without the extra imaginary factor (which changes sign undercomplex conjugation) the two sides of the equation differ by a minus sign.Making the differential operator Hermitian does not change in any essentialway its commutation properties, <strong>and</strong> the commutation relation of the two basicquantum operators reads[[ p x ,x] = −i ∂ ]∂x ,x = −i. (19.23)Corresponding results hold when x is replaced, in both operators, by y or z.However, it should be noted that if different Cartesian coordinates appear in thetwo operators then the operators commute, i.e.[ p x ,y] = [ p x ,z] = [ p y ,x ] = [ p y ,z ] = [ p z ,x] = [ p z ,y] =0.(19.24)As an illustration of the substitution rules, we now construct the Hamiltonian(the quantum-mechanical energy operator) H <strong>for</strong> a particle of mass m movingin a potential V (x, y, z) when it has one of its allowed energy values, i.e itsenergy is E n ,whereH|ψ n 〉 = E n |ψ n 〉. This latter equation when expressed in aparticular coordinate system is the Schrödinger equation <strong>for</strong> the particle. In termsof position <strong>and</strong> momentum, the total classical energy of the particle is given byE = p22m + V (x, y, z) =p2 x + p 2 y + p 2 z+ V (x, y, z).2mSubstituting −i∂/∂x <strong>for</strong> p x (<strong>and</strong> similarly <strong>for</strong> p y <strong>and</strong> p z ) in the first term on the657


QUANTUM OPERATORSRHS gives(−i) 2 ∂ ∂2m ∂x ∂x + (−i)2 ∂ ∂2m ∂y ∂y + (−i)2 ∂ ∂2m ∂z ∂z .The potential energy V , being a function of position only, becomes a purelymultiplicative operator, thus creating the full expression <strong>for</strong> the Hamiltonian,( ∂2H = − 22m∂x 2 + ∂2∂y 2 + ∂2∂z 2 )+ V (x, y, z),<strong>and</strong> giving the corresponding Schrödinger equation as(Hψ n = − 2 ∂ 2 )ψ n2m ∂x 2 + ∂2 ψ n∂y 2 + ∂2 ψ n∂z 2 + V (x, y, z)ψ n = E n ψ n .We are not so much concerned in this section with solving such differentialequations, but with the commutation properties of the operators from which theyare constructed. To this end, we now turn our attention to the topic of angularmomentum, the operators <strong>for</strong> which can be constructed in a straight<strong>for</strong>wardmanner from the two basic sets.19.2.1 Angular momentum operatorsAs required by the substitution rules, we start by expressing angular momentumin terms of the classical quantities r <strong>and</strong> p, namely L = r × p with CartesiancomponentsL z = xp y − yp x , L x = yp z − zp y , L y = zp x − xp z .Making the substitutions (19.22) yields as the corresponding quantum-mechanicaloperators(L z = −i x ∂∂y − y ∂ ),∂x(L x = −i y ∂ ∂z − z ∂ ), (19.25)∂y(L y = −i z ∂∂x − x ∂ ).∂zIt should be noted that <strong>for</strong> xp y , say, x <strong>and</strong> ∂/∂y commute, <strong>and</strong> there is noambiguity about the way it is to be carried into its quantum <strong>for</strong>m. Further, sincethe operators corresponding to each of its factors commute <strong>and</strong> are Hermitian,the operator corresponding to the product is Hermitian. This was shown directly<strong>for</strong> matrices in exercise 8.7, <strong>and</strong> can be verified using equation (17.16).The first question that arises is whether or not these three operators commute.658


19.2 PHYSICAL EXAMPLES OF OPERATORSConsider firstNow consider(L x L y = − 2 y ∂ ∂z − z ∂ )(z ∂∂y ∂x − x ∂ )∂z(= − 2 y ∂ ∂2 ∂2+ yz − yx∂x ∂z∂x ∂z 2 − ∂ 2z2 ∂2+ zx∂y∂x ∂y∂z(L y L x = − 2 z ∂∂x − x ∂ )(y ∂ ∂z ∂z − z ∂ )∂y(= − 2 zy ∂2∂x∂z − ∂ 2z2 ∂2− xy∂x∂y ∂z 2 + x ∂ ∂2+ xz∂y ∂z∂yThese two expressions are not the same. The difference between them, i.e. thecommutator of L x <strong>and</strong> L y , is given by[ ]Lx ,L y = Lx L y − L y L x = (x 2 ∂∂y − y ∂ )= iL z . (19.26)∂xThis, <strong>and</strong> two similar results obtained by permutting x, y <strong>and</strong> z cyclically,summarise the commutation relationships between the quantum operators correspondingto the three Cartesian components of angular momentum:[ ]Lx ,L y = iLz ,[ ]Ly ,L z = iLx , (19.27)[ L z ,L x ] = iL y .As well as its separate components of angular momentum, the total angularmomentum associated with a particular state |ψ〉 is a physical quantity of interest.This is measured by the operator corresponding to the sum of squares of itscomponents,).).L 2 = L 2 x + L 2 y + L 2 z. (19.28)This is an Hermitian operator, as each term in it is the product of two Hermitianoperators that (trivially) commute. It might seem natural to want to ‘take thesquare root’ of this operator, but such a process is undefined <strong>and</strong> we will notpursue the matter.We next show that, although no two of its components commute, the totalangular momentum operator does commute with each of its components. In theproof we use some of the properties (19.17) to (19.20) <strong>and</strong> result (19.27). We begin659


QUANTUM OPERATORSwith[L 2 ,L z]=[L2x + L 2 y + L 2 z,L z]= L x [ L x ,L z ] + [ L x ,L z ] L x[ ] [ ]+ L y Ly ,L z + Ly ,L z Ly + [ L 2 ]z,L z= L x (−i)L y +(−i)L y L x + L y (i)L x +(i)L x L y +0=0.Thus operators L 2 <strong>and</strong> L z commute <strong>and</strong>, continuing in the same way, it can beshown that[L 2 ,L x]=[L 2 ,L y]=[L 2 ,L z]=0. (19.29)Eigenvalues of the angular momentum operatorsWe will now use the commutation relations <strong>for</strong> L 2 <strong>and</strong> its components to findthe eigenvalues of L 2 <strong>and</strong> L z , without reference to any specific wavefunction. Inother words, the eigenvalues of the operators follow from the structure of theircommutators. There is nothing particular about L z ,<strong>and</strong>L x or L y could equallywell have been chosen, though, in general, it is not possible to find states that aresimultaneously eigenstates of two or more of L x , L y <strong>and</strong> L z .To help with the calculation, it is convenient to define the two operatorsU ≡ L x + iL y <strong>and</strong> D ≡ L x − iL y .These operators are not Hermitian; they are in fact Hermitian conjugates, in thatU † = D <strong>and</strong> D † = U, but they do not represent measurable physical quantities.We first note their multiplication <strong>and</strong> commutation properties:UD =(L x + iL y )(L x − iL y )=L 2 x + L 2 y + i [ ]L y ,L x= L 2 − L 2 z + L z , (19.30)DU =(L x − iL y )(L x + iL y )=L 2 x + L2 y − i [ ]L y ,L x= L 2 − L 2 z − L z , (19.31)[ L z ,U] = [ L z ,L x ] + i [ ]L z ,L y = iLy + L x = U, (19.32)[ L z ,D] = [ L z ,L x ] − i [ ]L z ,L y = iLy − L x = −D. (19.33)In the same way as was shown <strong>for</strong> matrices, it can be demonstrated that if twooperators commute they have a common set of eigenstates. Since L 2 <strong>and</strong> L zcommute they possess such a set; let one of the set be |ψ〉 withL 2 |ψ〉 = a|ψ〉 <strong>and</strong> L z |ψ〉 = b|ψ〉.Now consider the state |ψ ′ 〉 = U|ψ〉 <strong>and</strong> the actions of L 2 <strong>and</strong> L z upon it.660


19.2 PHYSICAL EXAMPLES OF OPERATORSConsider first L 2 |ψ ′ 〉, recalling that L 2 commutes with both L x <strong>and</strong> L y <strong>and</strong> hencewith U:L 2 |ψ ′ 〉 = L 2 U|ψ〉 = UL 2 |ψ〉 = Ua|ψ〉 = aU|ψ〉 = a|ψ ′ 〉.Thus, |ψ ′ 〉 is also an eigenstate of L 2 , corresponding to the same eigenvalue as|ψ〉. Now consider the action of L z :L z |ψ ′ 〉 = L z U|ψ〉=(UL z + U)|ψ〉, using[ L z ,U] = U,= Ub|ψ〉 + U|ψ〉=(b + )U|ψ〉=(b + )|ψ ′ 〉.Thus, |ψ ′ 〉 is also an eigenstate of L z , but with eigenvalue b + .In summary, the effect of U acting upon |ψ〉 is to produce a new state thathas the same eigenvalue <strong>for</strong> L 2 <strong>and</strong> is still an eigenstate of L z , though with thateigenvalue increased by . An exactly analogous calculation shows that the effectof D acting upon |ψ〉 is to produce another new state, one that also has the sameeigenvalue <strong>for</strong> L 2 <strong>and</strong> is also still an eigenstate of L z , though with the eigenvaluedecreased by in this case. For these reasons, U <strong>and</strong> D are usually known asladder operators.It is clear that, by starting from any arbitrary eigenstate <strong>and</strong> repeatedly applyingeither U or D, we could generate a series of eigenstates, all of which have theeigenvalue a <strong>for</strong> L 2 , but increment in their L z eigenvalues by ±. However, wealso have the physical requirement that, <strong>for</strong> real values of the z-component, itssquare cannot exceed the square of the total angular momentum, i.e. b 2 ≤ a. Thusb has a maximum value c that satisfiesc 2 ≤ a but (c + ) 2 >a;let the corresponding eigenstate be |ψ u 〉 with L z |ψ u 〉 = c|ψ u 〉. Now it is still truethatL z U|ψ u 〉 =(c + )U|ψ u 〉,<strong>and</strong>, to make this compatible with the physical constraint, we must have thatU|ψ u 〉 is the zero ket vector |∅〉. Now, using result (19.31), we haveDU|ψ u 〉 =(L 2 − L 2 z − L z )|ψ u 〉,⇒ 0|∅〉= D|∅〉=(a 2 − c 2 − c)|ψ u 〉,⇒ a = c(c + ).This gives the relationship between a <strong>and</strong> c. We now establish the possible <strong>for</strong>ms<strong>for</strong> c.If we start with eigenstate |ψ u 〉, which has the highest eigenvalue c <strong>for</strong> L z ,<strong>and</strong>661


QUANTUM OPERATORSoperate repeatedly on it with the (down) ladder operator D, we will generate astate |ψ d 〉 which, whilst still an eigenstate of L 2 with eigenvalue a, hasthelowestphysically possible value, d say, <strong>for</strong> the eigenvalue of L z . If this happens after noperations we will have that d = c − n <strong>and</strong>L z |ψ d 〉 =(c − n)|ψ d 〉.Arguing in the same way as previously that D|ψ d 〉 must be an unphysical ketvector, we conclude that0|∅〉= U|∅〉= UD|ψ d 〉=(L 2 − L 2 z + L z )|ψ d 〉, using (19.30),=[a − (c − n) 2 + (c − n)]|ψ d 〉⇒ a =(c − n) 2 − (c − n).Equating the two results <strong>for</strong> a givesc 2 + c = c 2 − 2cn + n 2 2 − c + n 2 ,2c(n +1)=n(n +1),c = 1 2 n.Since n is necessarily integral, c is an integer multiple of 1 2. This result is validirrespective of which eigenstate |ψ〉 we started with, though the actual value ofthe integer n depends on |ψ u 〉 <strong>and</strong> hence upon |ψ〉.Denoting 1 2n by l we can say that the possible eigenvalues of the operatorL z , <strong>and</strong> hence the possible results of a measurement of the z-component of theangular momentum of a system, are given byl, (l − 1), (l − 2), ... , −l.The value of a <strong>for</strong> all 2l + 1 of the corresponding states,|ψ u 〉, D|ψ u 〉, D 2 |ψ u 〉, ... , D 2l |ψ u 〉,is l(l +1) 2 .The similarity of <strong>for</strong>m between this eigenvalue <strong>and</strong> that appearing in Legendre’sequation is not an accident. It is intimately connected with the facts (i) that L 2is a measure of the rotational kinetic energy of a particle in a system centredon the origin, <strong>and</strong> (ii) that in spherical polar coordinates L 2 has the same <strong>for</strong>mas the angle-dependent part of ∇ 2 , which, as we have seen, is itself proportionalto the quantum-mechanical kinetic energy operator. Legendre’s equation <strong>and</strong>the associated Legendre equation arise naturally when ∇ 2 ψ = f(r) issolvedinspherical polar coordinates using the method of separation of variables discussedin chapter 21.The derivation of the eigenvalues l(l+1) 2 <strong>and</strong> m, with −l ≤ m ≤ l, dependsonly on the commutation relationships between the corresponding operators. Any662


19.2 PHYSICAL EXAMPLES OF OPERATORSother set of four operators with the same commutation structure would resultin the same eigenvalue spectrum. In fact, quantum mechanically, orbital angularmomentum is restricted to cases in which n is even <strong>and</strong> so l is an integer; thisis in accord with the requirement placed on l if solutions to ∇ 2 ψ = f(r) thatarefinite on the polar axis are to be obtained. The non-classical notion of internalangular momentum (spin) <strong>for</strong> a particle provides a set of operators that are ableto take both integral <strong>and</strong> half-integral multiples of as their eigenvalues.We have already seen that, <strong>for</strong> a state |l, m〉 that has a z-component ofangular momentum m, the state U|l, m〉 is one with its z-component of angularmomentum equal to (m +1). But the new state ket vector so produced is notnecessarily normalised so as to make 〈l, m +1| l, m +1〉 = 1. We will concludethis discussion of angular momentum by calculating the coefficients µ m <strong>and</strong> ν m inthe equationsU|l, m〉 = µ m |l, m +1〉 <strong>and</strong> D|l, m〉 = ν m |l, m − 1〉on the basis that 〈l, r | l, r〉 = 1 <strong>for</strong> all l <strong>and</strong> r.To do so, we consider the inner product I = 〈l, m |DU| l, m〉, evaluated in twodifferent ways. We have already noted that U <strong>and</strong> D are Hermitian conjugates<strong>and</strong> so I canbewrittenasI = 〈l, m |U † U| l, m〉 = µ ∗ m〈l, m | l, m〉µ m = |µ m | 2 .But, using equation (19.31), it can also be expressed asThus we are required to haveI = 〈l, m |L 2 − L 2 z − L z | l, m〉= 〈l, m |l(l +1) 2 − m 2 2 − m 2 | l, m〉=[l(l +1) 2 − m 2 2 − m 2 ] 〈l, m | l, m〉=[l(l +1)− m(m +1)] 2 .|µ m | 2 =[l(l +1)− m(m +1)] 2 ,but can choose that all µ m are real <strong>and</strong> non-negative (recall that |m| ≤l). Asimilar calculation can be used to calculate ν m . The results are summarised in theequationsU| l, m〉 = √ l(l +1)− m(m +1)| l, m +1〉, (19.34)D| l, m〉 = √ l(l +1)− m(m − 1) | l, m − 1〉. (19.35)It can easily be checked that U|l, l〉 = |∅〉= D|l, −l〉.663


QUANTUM OPERATORS19.2.2 Uncertainty principlesThe next topic we explore is the quantitative consequences of a non-zero commutator<strong>for</strong> two quantum (Hermitian) operators that correspond to physicalvariables.As previously noted, the expectation value in a state |ψ〉 of the physical quantityA corresponding to the operator A is E[A] =〈ψ| A |ψ〉. Any one measurement ofA can only yield one of the eigenvalues of A. But if repeated measurements couldbe made on a large number of identical systems, a discrete or continuous rangeof values would be obtained. It is a natural extension of normal data analysisto measure the uncertainty in the value of A by the observed variance in themeasured values of A, denoted by (∆A) 2 <strong>and</strong> calculated as the average value of(A − E[A]) 2 . The expected value of this variance <strong>for</strong> the state |ψ〉 is given by〈ψ |(A − E[A]) 2 |ψ〉.We now give a mathematical proof that there is a theoretical lower limit <strong>for</strong>the product of the uncertainties in any two physical quantities, <strong>and</strong> we start byproving a result similar to the Schwarz inequality. Let |u〉 <strong>and</strong> |v〉 be any two statevectors <strong>and</strong> let λ be any real scalar. Then consider the vector |w〉 = |u〉 + λ|v〉 <strong>and</strong>,in particular, note that0 ≤〈w | w〉 = 〈u | u〉 + λ(〈u | v〉 + 〈v | u〉)+λ 2 〈v | v〉.This is a quadratic inequality in λ <strong>and</strong> there<strong>for</strong>e the quadratic equation <strong>for</strong>medby equating the RHS to zero must have no real roots. The coefficient of λ is(〈u | v〉 + 〈v | u〉) =2Re〈u | v〉 <strong>and</strong> its square is thus ≥ 0. The condition <strong>for</strong> no realroots of the quadratic is there<strong>for</strong>e0 ≤ (〈u | v〉 + 〈v | u〉) 2 ≤ 4〈u | u〉〈v | v〉. (19.36)This result will now be applied to state vectors constructed from |ψ〉, the statevector of the particular system <strong>for</strong> which we wish to establish a relationship betweenthe uncertainties in the two physical variables corresponding to (Hermitian)operators A <strong>and</strong> B. WetakeThenFurther,|u〉 =(A − E[A]) | ψ〉 <strong>and</strong> |v〉 = i(B − E[B]) | ψ〉. (19.37)〈u | u〉 = 〈ψ |(A − E[A]) 2 |ψ〉 =(∆A) 2 ,〈v | v〉 = 〈ψ |(B − E[B]) 2 |ψ〉 =(∆B) 2 .〈u | v〉 = 〈ψ | (A − E[A])i(B − E[B]) | ψ〉= i〈ψ | AB | ψ〉−iE[A]〈ψ | B | ψ〉−iE[B]〈ψ | A | ψ〉 + iE[A]E[B]〈ψ | ψ〉= i〈ψ | AB | ψ〉−iE[A]E[B].664


19.2 PHYSICAL EXAMPLES OF OPERATORSIn the second line, we have moved expectation values, which are purely numbers,out of the inner products <strong>and</strong> used the normalisation condition 〈ψ|ψ〉 = 1.Similarly〈v | u〉 = −i〈ψ | BA| ψ〉 + iE[A]E[B].Adding these two results gives<strong>and</strong> substitution into (19.36) yields〈u | v〉 + 〈v | u〉 = i〈ψ | AB − BA|ψ〉,0 ≤ (i〈ψ | AB − BA|ψ〉) 2 ≤ 4(∆A) 2 (∆B) 2At first sight, the middle term of this inequality might appear to be negative, butthis is not so. Since A <strong>and</strong> B are Hermitian, AB −BA is anti-Hermitian, as is easilydemonstrated. Since i is also anti-Hermitian, the quantity in the parentheses inthe middle term is real <strong>and</strong> its square non-negative. Rearranging the equation <strong>and</strong>expressing it in terms of the commutator of A <strong>and</strong> B gives the generalised <strong>for</strong>mof the Uncertainty Principle. For any particular state |ψ〉 of a system, this providesthe quantitative relationship between the minimum value that the product of theuncertainties in A <strong>and</strong> B can have <strong>and</strong> the expectation value, in that state, oftheir commutator,Immediate observations include the following:(∆A) 2 (∆B) 2 ≥ 1 4 |〈ψ | [ A, B ] | ψ〉|2 . (19.38)(i) If A <strong>and</strong> B commute there is no absolute restriction on the accuracy withwhich the corresponding physical quantities may be known. That is not tosay that ∆A <strong>and</strong> ∆B will always be zero, only that they may be.(ii) If the commutator of A <strong>and</strong> B is a constant, k ≠ 0, then the RHS ofequation (19.38) is necessarily equal to 1 4 |k|2 , whatever the <strong>for</strong>m of |ψ〉,<strong>and</strong> it is not possible to have ∆A =∆B =0.(iii) Since the RHS depends upon |ψ〉, it is possible, even <strong>for</strong> two operatorsthat do not commute, <strong>for</strong> the lower limit of (∆A) 2 (∆B) 2 to be zero. Thiswill occur if the commutator [ A, B ] is itself an operator whose expectationvalue in the particular state |ψ〉 happens to be zero.To illustrate the third of these, we might consider the components of angularmomentum discussed in the previous subsection. There, in equation (19.27), wefound that the commutator of the operators corresponding to the x- <strong>and</strong> y-components of angular momentum is non-zero; in fact, it has the value iL z .This means that if the state |ψ〉 of a system happened to be such that 〈ψ|L z |ψ〉 =0,as it would if, <strong>for</strong> example, it were the eigenstate of L z , |ψ〉 = |l, 0〉, then therewould be no fundamental reason why the physical values of both L x <strong>and</strong> L yshould not be known exactly. Indeed, if the state were spherically symmetric, <strong>and</strong>665


QUANTUM OPERATORShence <strong>for</strong>mally an eigenstate of L 2 with l = 0, all three components of angularmomentum could be (<strong>and</strong> are) known to be zero.◮Working in one dimension, show that the minimum value of the product ∆p x × ∆x <strong>for</strong>aparticleis 1 . Find the <strong>for</strong>m of the wavefunction that attains this minimum value <strong>for</strong> a2particle whose expectation values <strong>for</strong> position <strong>and</strong> momentum are ¯x <strong>and</strong> ¯p, respectively.We have already seen, in (19.23) that the commutator of p x <strong>and</strong> x is −i, aconstant.There<strong>for</strong>e, irrespective of the actual <strong>for</strong>m of |ψ〉, the RHS of (19.38) is 1 4 2 (see observation(ii) above). Thus, since all quantities are positive, taking the square roots of both sides ofthe equation shows directly that∆p x × ∆x ≥ 1 . 2Returning to the derivation of the Uncertainty Principle, we see that the inequality becomesan equality only when(〈u | v〉 + 〈v | u〉) 2 =4〈u | u〉〈v | v〉.The RHS of this equality has the value 4||u|| 2 ||v|| 2 <strong>and</strong> so, by virtue of Schwarz’s inequality,we have4‖u‖ 2 ‖v‖ 2 =(〈u | v〉 + 〈v | u〉) 2≤ (|〈u | v〉| + |〈v | u〉|) 2≤ (‖u‖‖v|| + ‖v‖‖u‖) 2=4‖u‖ 2 ‖v‖ 2 .Since the LHS is less than or equal to something that has the same value as itself, all ofthe inequalities are, in fact, equalities. Thus 〈u|v〉 = ‖u‖‖v‖, showing that |u〉 <strong>and</strong> |v〉 areparallel vectors, i.e. |u〉 = µ|v〉 <strong>for</strong> some scalar µ.We now trans<strong>for</strong>m this condition into a constraint that the wavefunction ψ = ψ(x)must satisfy. Recalling the definitions (19.37) of |u〉 <strong>and</strong> |v〉 in terms of |ψ〉, we have(−i d )dx − ¯p ψ = µi(x − ¯x)ψ,dψ[ µ(x − ¯x) − i¯p ]ψ =0.dx + 1 [ µ(x − ¯x)2The IF <strong>for</strong> this equation is exp− i¯px ], giving2 { [d µ(x − ¯x)2ψ expdx2− i¯px ]}=0,which, in turn, leads to[ ] ( )µ(x − ¯x)2 i¯pxψ(x) =A exp − exp .2From this it is apparent that the minimum uncertainty product ∆p x × ∆x is obtained whenthe probability density |ψ(x)| 2 has the <strong>for</strong>m of a Gaussian distribution centred on ¯x. Thevalue of µ is not fixed by this consideration <strong>and</strong> it could be anything (positive); a largevalue <strong>for</strong> µ would yield a small value <strong>for</strong> ∆x but a correspondingly large one <strong>for</strong> ∆p x . ◭666


19.2 PHYSICAL EXAMPLES OF OPERATORS19.2.3 Annihilation <strong>and</strong> creation operatorsAs a final illustration of the use of operator methods in physics we consider theirapplication to the quantum mechanics of a simple harmonic oscillator (s.h.o.).Although we will start with the conventional description of a one-dimensionaloscillator, using its position <strong>and</strong> momentum, we will recast the description interms of two operators <strong>and</strong> their commutator <strong>and</strong> show that many importantconclusions can be reached from studying these alone.The Hamiltonian <strong>for</strong> a particle of mass m with momentum p moving in aone-dimensional parabolic potential V (x) = 1 2 kx2 isH = p22m + 1 2 kx2 = p22m + 1 2 mω2 x 2 ,where its classical frequency of oscillation ω is given by ω 2 = k/m. We recall thatthe corresponding operators, p <strong>and</strong> x, do not commute <strong>and</strong> that [ p, x ] = −i.In analogy with the ladder operators used when discussing angular momentum,we define two new operators:√ mω2 x + ip √2mω<strong>and</strong> A † ≡√ mω2 x − ip √2mω.A ≡(19.39)Since both x <strong>and</strong> p are Hermitian, A <strong>and</strong> A † are Hermitian conjugates, thoughneither is Hermitian <strong>and</strong> they do not represent physical quantities that can bemeasured.Now consider the two products A † A <strong>and</strong> AA † :A † A = mω 2x2 − ipx2 + ixp2 + p22mω = H ω − i 2 [ p, x ] = H ω − 2 ,AA † = mω 2x2 + ipx2 − ixp2 + p22mω = H ω + i 2 [ p, x ] = H ω + 2 .From these it follows that<strong>and</strong> thatFurther,Similarly,H = 1 2 ω(A† A + AA † ) (19.40)[A, A† ] = . (19.41)[ H,A] = [ 12 ω(A† A + AA † ),A ]= 1 2 ω ( A † 0+ [ A † ,A ] A + A [ A † ,A ] +0A †)= 1 2ω(−A − A) =−ωA. (19.42)[H,A† ] = ωA † (19.43).Be<strong>for</strong>e we apply these relationships to the question of the energy spectrum ofthe s.h.o., we need to prove one further result. This is that if B is an Hermitianoperator then 〈ψ | B 2 | ψ〉 ≥0<strong>for</strong>any |ψ〉. The proof, which involves introducing667


QUANTUM OPERATORSan arbitrary complete set of orthonormal base states |φ i 〉 <strong>and</strong> using equation(19.11), is as follows:〈ψ | B 2 | ψ〉 = 〈ψ | B × 1 × B | ψ 〉= ∑ i= ∑ i= ∑ i= ∑ i= ∑ i〈ψ | B |φ i 〉〈φ i | B |ψ〉〈ψ | B |φ i 〉 ( 〈φ i | B |ψ〉 ∗) ∗〈ψ | B |φ i 〉 ( 〈ψ | B † |φ i 〉 ) ∗〈ψ | B |φ i 〉〈ψ | B |φ i 〉 ∗ , since B is Hermitian,|〈ψ | B |φ i 〉| 2 ≥ 0.We note, <strong>for</strong> future reference, that the Hamiltonian H <strong>for</strong> the s.h.o. is the sum oftwo terms each of this <strong>for</strong>m <strong>and</strong> there<strong>for</strong>e conclude that 〈ψ|H|ψ〉 ≥0 <strong>for</strong> all |ψ〉.The energy spectrum of the simple harmonic oscillatorLet the normalised ket vector |n〉 (or |E n 〉) denote the nth energy state of the s.h.o.with energy E n . Then it must be an eigenstate of the (Hermitian) Hamiltonian H<strong>and</strong> satisfyH|n〉 = E n |n〉 with 〈m|n〉 = δ mn .Now consider the state A|n〉 <strong>and</strong> the effect of H upon it:HA|n〉 = AH|n〉−ωA|n〉, using (19.42),= AE n |n〉−ωA|n〉=(E n − ω)A|n〉.Thus A|n〉 is an eigenstate of H corresponding to energy E n − ω <strong>and</strong> must besome multiple of the normalised ket vector |E n − ω〉, i.e.A| E n 〉≡A|n〉 = c n |E n − ω〉,where c n is not necessarily of unit modulus. Clearly, A is an operator thatgenerates a new state that is lower in energy by ω; it can thus be compared tothe operator D, which has a similar effect in the context of the z-component ofangular momentum. Because it possesses the property of reducing the energy ofthe state by ω, which, as we will see, is one quantum of excitation energy <strong>for</strong> theoscillator, the operator A is called an annihilation operator. Repeated applicationof A, m times say, will produce a state whose energy is mω lower than that ofthe original:A m |E n 〉 = c n c n−1 ···c n−m+1 |E n − mω〉. (19.44)668


19.2 PHYSICAL EXAMPLES OF OPERATORSIn a similar way it can be shown that A † parallels the operator U of our angularmomentum discussion <strong>and</strong> creates an additional quantum of energy each time itis applied:(A † ) m |E n 〉 = d n d n+1 ···d n+m−1 |E n + mω〉. (19.45)It is there<strong>for</strong>e known as a creation operator.As noted earlier, the expectation value of the oscillator’s energy operator〈ψ|H|ψ〉 must be non-negative, <strong>and</strong> there<strong>for</strong>e it must have a lowest value. Letthis be E 0 , with corresponding eigenstate |0〉. Since the energy-lowering propertyof A applies to any eigenstate of H, in order to avoid a contradiction we musthave that A|0〉 = |∅〉. It then follows from (19.40) thatH|0〉 = 1 2 ω(A† A + AA † )|0〉= 1 2 ωA† A|0〉 + 1 2 ω(A† A + )|0〉, using (19.41),=0+0+ 1 2ω|0〉. (19.46)This shows that the commutator structure of the operators <strong>and</strong> the <strong>for</strong>m of theHamiltonian imply that the lowest energy (its ground-state energy) is 1 2ω; thisis a result that has been derived without explicit reference to the correspondingwavefunction. This non-zero lowest value <strong>for</strong> the energy, known as the zero-pointenergy of the oscillator, <strong>and</strong> the discrete values <strong>for</strong> the allowed energy statesare quantum-mechanical in origin; classically such an oscillator could have anynon-negative energy, including zero.Working back from this result, we see that the energy levels of the s.h.o. are12 ω, 32 ω, 52 ω,... ,(m + 1 2)ω,..., <strong>and</strong> that the corresponding (unnormalised)ket vectors can be written as|0〉, A † |0〉, (A † ) 2 |0〉, ... , (A † ) m |0〉, ... .This notation, <strong>and</strong> elaborations of it, are often used in the quantum treatment ofclassical fields such as the electromagnetic field. Thus, as the reader should verify,A(A † ) 3 A 2 A † A(A † ) 4 |0〉 is a state with energy 9 2 ω, whilst A(A† ) 3 A 5 A † A(A † ) 4 |0〉 isnot a physical state at all.The normalisation of the eigenstatesIn order to make quantitative calculations using the previous results we need toestablish the values of the c n <strong>and</strong> d n that appear in equations (19.44) <strong>and</strong> (19.45).To do this, we first establish the operator recurrence relationA m (A † ) m = A m−1 (A † ) m A + mA m−1 (A † ) m−1 . (19.47)669


QUANTUM OPERATORSThe proof, which makes repeated use of [ A, A † ] = , is as follows:A m (A † ) m = A m−1 AA † (A † ) m−1= A m−1 (A † A + )(A † ) m−1= A m−1 A † A(A † ) m−1 + A m−1 (A † ) m−1= A m−1 A † (A † A + )(A † ) m−2 + A m−1 (A † ) m−1= A m−1 (A † ) 2 A(A † ) m−2 + A m−1 A † (A † ) m−2 + A m−1 (A † ) m−1= A m−1 (A † ) 2 (A † A + )(A † ) m−3 +2A m−1 (A † ) m−1..= A m−1 (A † ) m A + mA m−1 (A † ) m−1 .Now we take the expectation values in the ground state |0〉 of both sides ofthis operator equation <strong>and</strong> note that the first term on the RHS is zero since itcontains the term A|0〉. The non-vanishing terms are〈0 | A m (A † ) m | 0〉 = m〈0 | A m−1 (A † ) m−1 | 0〉.The LHS is the square of the norm of (A † ) m |0〉, <strong>and</strong>, from equation (19.45), it isequal to|d 0 | 2 |d 1 | 2 ···|d m−1 | 2 〈0 | 0〉.Similarly, the RHS is equal tom |d 0 | 2 |d 1 | 2 ···|d m−2 | 2 〈0 | 0〉.It follows that |d m−1 | 2 = m <strong>and</strong>, taking all coefficients as real, d m = √ (m +1).Thus the correctly normalised state of energy (n + 1 2), obtained by repeatedapplication of A † to the ground state, is given by|n〉 =(A† ) n|0〉. (19.48)(n! n )1/2To evaluate the c n , we note that, from the commutator of A <strong>and</strong> A † ,[A, A† ] |n〉 = AA † |n〉−A † A|n〉 |n〉 = √ (n +1) A |n +1〉−c n A † |n − 1〉= √ √(n +1) c n+1 |n〉−c n n |n〉, = √ √(n +1) c n+1 − c n n,which has the obvious solution c n = √ n. To summarise:c n = √ n <strong>and</strong> d n = √ (n +1). (19.49)We end this chapter with another worked example. This one illustrates howthe operator <strong>for</strong>malism that we have developed can be used to obtain results670


19.3 EXERCISESthat would involve a number of non-trivial integrals if tackled using explicitwavefunctions.◮Given that the first-order change in the ground-state energy of a quantum system whenit is perturbed by a small additional term H ′ in the Hamiltonian is 〈0|H ′ |0〉, find the firstorderchange in the energy of a simple harmonic oscillator in the presence of an additionalpotential V ′ (x) =λx 3 + µx 4 .From the definitions of A <strong>and</strong> A † , equation (19.39), we can writex = √ 1 (A + A † ) ⇒ H ′ λ=2mω (2mω) (A + 3/2 A† ) 3 µ+(2mω) (A + 2 A† ) 4 .We now compute successive values of (A + A † ) n |0〉 <strong>for</strong> n =1, 2, 3, 4, remembering thatA |n〉 = √ n |n − 1〉 <strong>and</strong> A † |n〉 = √ (n +1) |n +1〉 :(A + A † ) |0〉 =0+ 1/2 |1〉,(A + A † ) 2 |0〉 = |0〉 + √ 2 |2〉,(A + A † ) 3 |0〉 =0+ 3/2 |1〉 +2 3/2 |1〉 + √ 6 3/2 |3〉=3 3/2 |1〉 + √ 6 3/2 |3〉,(A + A † ) 4 |0〉 =3 2 |0〉 + √ 18 2 |2〉 + √ 18 2 |2〉 + √ 24 2 |4〉.To find the energy shift we need to <strong>for</strong>m the inner product of each of these state vectorswith |0〉. But|0〉 is orthogonal to all |n〉 if n ≠ 0. Consequently, the term 〈0| (A + A † ) 3 |0〉in the expectation value is zero, <strong>and</strong> in the expression <strong>for</strong> 〈0| (A + A † ) 4 |0〉 only the firstterm is non-zero; its value is 3 2 . The perturbation energy is thus given by〈0 | H ′ | 0〉 = 3µ2(2mω) . 2It could have been anticipated on symmetry grounds that the expectation of λx 3 , anodd function of x, would be zero, but the calculation gives this result automatically. Thecontribution of the quadratic term in the perturbation would have been much harder toanticipate! ◭19.3 Exercises19.1 Show that the commutator of two operators that correspond to two physicalobservables cannot itself correspond to another physical observable.19.2 By expressing the operator L z , corresponding to the z-component of angularmomentum, in spherical polar coordinates (r, θ, φ), show that the angular momentumof a particle about the polar axis cannot be known at the same time asits azimuthal position around that axis.19.3 In quantum mechanics, the time dependence of the state function |ψ〉 of a systemis given, as a further postulate, by the equationi ∂ |ψ〉 = H|ψ〉,∂twhere H is the Hamiltonian of the system. Use this to find the time dependenceof the expectation value 〈A〉 of an operator A that itself has no explicit timedependence. Hence show that operators that commute with the Hamiltoniancorrespond to the classical ‘constants of the motion’.671


QUANTUM OPERATORSFor a particle of mass m moving in a one-dimensional potential V (x), proveEhrenfest’s theorem:〈 〉d〈p x 〉 dVd〈x〉= −<strong>and</strong> = 〈p x〉dt dxdt m .19.4 Show that the Pauli matricesS x = 1 2 (0 11 0), S y = 1 2 (0 −ii 0)( )1 0, S z = 1 ,2 0 −1which are used as the operators corresponding to intrinsic spin of 1 in nonrelativisticquantum mechanics, satisfy S 2 x = S 2 y = S 2 z = 1 4 2 I, <strong>and</strong> have the2same commutation properties as the components of orbital angular momentum.Deduce that any state |ψ〉 represented by the column vector (a, b) T is an eigenstateof S 2 with eigenvalue 3 2 /4.19.5 Find closed-<strong>for</strong>m expressions <strong>for</strong> cos C <strong>and</strong> sin C, whereC is the matrixC =(1 11 −1Demonstrate that the ‘expected’ relationshipscos 2 C +sin 2 C = I <strong>and</strong> sin 2C =2sinC cos Care valid.19.6 Operators A <strong>and</strong> B anticommute. Evaluate (A + B) 2n <strong>for</strong> a few values of n <strong>and</strong>hence propose an expression <strong>for</strong> c nr in the expansionn∑(A + B) 2n = c nr A 2n−2r B 2r .r=0Prove your proposed <strong>for</strong>mula <strong>for</strong> general values of n, using the method ofinduction.Show that∞∑ n∑cos(A + B) = d nr A 2n−2r B 2r ,n=0 r=0where the d nr are constants whose values ( you should ) determine.0 1By taking as A the matrix A =, confirm that your answer is1 0consistent with that obtained in exercise 19.5.19.7 Expressed in terms of the annihilation <strong>and</strong> creation operators A <strong>and</strong> A † discussedin the text, a system has an unperturbed Hamiltonian H 0 = ωA † A. The systemis disturbed by the addition of a perturbing Hamiltonian H 1 = gω(A + A † ),where g is real. Show that the effect of the perturbation is to move the wholeenergy spectrum of the system down by g 2 ω.19.8 For a system of N electrons in their ground state |0〉, the Hamiltonian isN∑ p 2 xH =n+ p 2 y n+ p 2 N∑z n+ V (x n ,y n ,z n ).2mn=1n=1Show that [ ]p 2 x n,x n = −2ipxn , <strong>and</strong> hence that the expectation value of thedouble commutator [[x, H ] ,x], wherex = ∑ Nn=1 x n,isgivenby〈0 | [[x, H ] ,x] | 0〉 = N2m .672).


19.3 EXERCISESNow evaluate the expectation value using the eigenvalue properties of H, namelyH|r〉 = E r |r〉, <strong>and</strong> deduce the sum rule <strong>for</strong> oscillation strengths,∞∑r=0(E r − E 0 )|〈r | x | 0〉| 2 = N22m .19.9 By considering the functionF(λ) =exp(λA)B exp(−λA),where A <strong>and</strong> B are linear operators <strong>and</strong> λ is a parameter, <strong>and</strong> finding itsderivatives with respect to λ, prove thate A Be −A = B + [ A, B ] + 1 2! [ A, [ A, B ]]+ 1 [ A, [ A, [ A, B ]]]+ ··· .3!Use this result to express( ) ( )iLx θ−iLx θexp L y expas a linear combination of the angular momentum operators L x , L y <strong>and</strong> L z .19.10 For a system containing more than one particle, the total angular momentum J<strong>and</strong> its components are represented by operators that have completely analogouscommutation relations to those <strong>for</strong> the operators <strong>for</strong> a single particle, i.e. J 2 haseigenvalue j(j +1) 2 <strong>and</strong> J z has eigenvalue m j <strong>for</strong> the state |j,m j 〉. The usualorthonormality relationship 〈j ′ ,m ′ j | j,m j〉 = δ j ′ j δ m ′jm jis also valid.A system consists of two (distinguishable) particles A <strong>and</strong> B. ParticleA is inan l = 3 state <strong>and</strong> can have state functions of the <strong>for</strong>m |A, 3,m A 〉, whilst B isin an l = 2 state with possible state functions |B,2,m B 〉. The range of possiblevalues <strong>for</strong> j is |3 − 2| ≤j ≤|3+2|, i.e.1≤ j ≤ 5, <strong>and</strong> the overall state functioncanbewrittenas|j,m j 〉 =∑m A m B| A, 3,m A 〉|B,2,m B 〉.m A +m B =m jC jm jThe numerical coefficients C jm jm A m Bare known as Clebsch–Gordon coefficients.Assume (as can be shown) that the ladder operators U(AB) <strong>and</strong>D(AB) <strong>for</strong>the system can be written as U(A) +U(B) <strong>and</strong>D(A) +D(B), respectively, <strong>and</strong>that they lead to relationships equivalent to (19.34) <strong>and</strong> (19.35) with l replacedby j <strong>and</strong> m by m j .(a) Apply the operators to the (obvious) relationship|AB, 5, 5〉 = |A, 3, 3〉|B,2, 2〉to show that√√6|AB, 5, 4〉 = |A, 3, 2〉|B,2, 2〉 + 4|A, 3, 3〉|B,2, 1〉.10 10(b) Find, to within an overall sign, the real coefficients c <strong>and</strong> d in the expansion|AB, 4, 4〉 = c|A, 3, 2〉|B,2, 2〉 + d|A, 3, 3〉|B,2, 1〉by requiring it to be orthogonal to |AB, 5, 4〉. Check your answer by consideringU(AB)|AB, 4, 4〉.(c) Find, to within an overall sign, <strong>and</strong> as efficiently as possible, an expression<strong>for</strong> |AB, 4, −3〉 as a sum of products of the <strong>for</strong>m |A, 3,m A 〉|B,2,m B 〉.673


QUANTUM OPERATORS19.4 Hints <strong>and</strong> answers19.1 Show that the commutator is anti-Hermitian.19.3 Use the Hermitian conjugate of the given equation to obtain the time dependenceof 〈ψ|. The rate of change of 〈ψ|A|ψ〉 is i〈ψ| [ H,A] |ψ〉. Note that [ H,p x ] =[ V,p x ] <strong>and</strong> [ H,x] = [ p 2 x,x ] /2m.19.5 Show that C 2 =2I.cos C =cos √ ( )1 02, sin C = sin √ ( )2 1 1√ .0 12 1 −119.7 Express the total Hamiltonian in terms of B = A + gI <strong>and</strong> determine the valueof [ B,B ] † .19.9 Show that, if F (n) is the nth derivative of F(λ), then F (n+1) = [ A, F ] (n) . Use a Taylorseries in λ to evaluate F(1), using derivatives evaluated at λ = 0. Successivelyreduce the level of nesting of each multiple commutator by using the result ofevaluating the previous term. The given expression reduces to cos θL y − sin θL z .674


20Partial differential equations:general <strong>and</strong> particular solutionsIn this chapter <strong>and</strong> the next the solution of differential equations of typestypically encountered in the physical sciences <strong>and</strong> engineering is extended tosituations involving more than one independent variable. A partial differentialequation (PDE) is an equation relating an unknown function (the dependentvariable) of two or more variables to its partial derivatives with respect tothose variables. The most commonly occurring independent variables are thosedescribing position <strong>and</strong> time, <strong>and</strong> so we will couch our discussion <strong>and</strong> examplesin notation appropriate to them.As in other chapters we will focus our attention on the equations that arisemost often in physical situations. We will restrict our discussion, there<strong>for</strong>e, tolinear PDEs, i.e. those of first degree in the dependent variable. Furthermore, wewill discuss primarily second-order equations. The solution of first-order PDEswill necessarily be involved in treating these, <strong>and</strong> some of the methods discussedcan be extended without difficulty to third- <strong>and</strong> higher-order equations. We shallalso see that many ideas developed <strong>for</strong> ordinary differential equations (ODEs)can be carried over directly into the study of PDEs.In this chapter we will concentrate on general solutions of PDEs in termsof arbitrary functions <strong>and</strong> the particular solutions that may be derived fromthem in the presence of boundary conditions. We also discuss the existence <strong>and</strong>uniqueness of the solutions to PDEs under given boundary conditions.In the next chapter the methods most commonly used in practice <strong>for</strong> obtainingsolutions to PDEs subject to given boundary conditions will be considered. Thesemethods include the separation of variables, integral trans<strong>for</strong>ms <strong>and</strong> Green’sfunctions. This division of material is rather arbitrary <strong>and</strong> has been made onlyto emphasise the general usefulness of the latter methods. In particular, it willbe readily apparent that some of the results of the present chapter are infact solutions in the <strong>for</strong>m of separated variables, but arrived at by a differentapproach.675


PDES: GENERAL AND PARTICULAR SOLUTIONS20.1 Important partial differential equationsMost of the important PDEs of physics are second-order <strong>and</strong> linear. In order togain familiarity with their general <strong>for</strong>m, some of the more important ones willnow be briefly discussed. These equations apply to a wide variety of differentphysical systems.Since, in general, the PDEs listed below describe three-dimensional situations,the independent variables are r <strong>and</strong> t, wherer is the position vector <strong>and</strong> t istime. The actual variables used to specify the position vector r are dictated by thecoordinate system in use. For example, in Cartesian coordinates the independentvariables of position are x, y <strong>and</strong> z, whereas in spherical polar coordinates theyare r, θ <strong>and</strong> φ. The equations may be written in a coordinate-independent manner,however, by the use of the Laplacian operator ∇ 2 .The wave equation20.1.1 The wave equation∇ 2 u = 1 ∂ 2 uc 2 ∂t 2 (20.1)describes as a function of position <strong>and</strong> time the displacement from equilibrium,u(r,t), of a vibrating string or membrane or a vibrating solid, gas or liquid. Theequation also occurs in electromagnetism, where u may be a component of theelectric or magnetic field in an elecromagnetic wave or the current or voltagealong a transmission line. The quantity c is the speed of propagation of the waves.◮ Find the equation satisfied by small transverse displacements u(x, t) of a uni<strong>for</strong>m string ofmass per unit length ρ held under a uni<strong>for</strong>m tension T , assuming that the string is initiallylocated along the x-axis in a Cartesian coordinate system.Figure 20.1 shows the <strong>for</strong>ces acting on an elemental length ∆s of the string. If the tensionT in the string is uni<strong>for</strong>m along its length then the net upward vertical <strong>for</strong>ce on theelement is∆F = T sin θ 2 − T sin θ 1 .Assuming that the angles θ 1 <strong>and</strong> θ 2 are both small, we may make the approximationsin θ ≈ tan θ. Since at any point on the string the slope tan θ = ∂u/∂x, the<strong>for</strong>cecanbewritten[ ∂u(x +∆x, t)∆F = T−∂x]∂u(x, t)∂x≈ T ∂2 u(x, t)∆x,∂x 2where we have used the definition of the partial derivative to simplify the RHS.This upward <strong>for</strong>ce may be equated, by Newton’s second law, to the product of themass of the element <strong>and</strong> its upward acceleration. The element has a mass ρ ∆s, whichisapproximately equal to ρ ∆x if the vibrations of the string are small, <strong>and</strong> so we haveρ ∆x ∂2 u(x, t)= T ∂2 u(x, t)∆x.∂t 2 ∂x 2676


20.1 IMPORTANT PARTIAL DIFFERENTIAL EQUATIONSuTθ 2∆sθ 1Txx +∆xxFigure 20.1tension T .The <strong>for</strong>ces acting on an element of a string under uni<strong>for</strong>mDividing both sides by ∆x we obtain, <strong>for</strong> the vibrations of the string, the one-dimensionalwave equation∂ 2 u∂x = 1 ∂ 2 u2 c 2 ∂t , 2where c 2 = T/ρ. ◭The longitudinal vibrations of an elastic rod obey a very similar equation tothat derived in the above example, namely∂ 2 u∂x 2 = ρ ∂ 2 uE ∂t 2 ;here ρ is the mass per unit volume <strong>and</strong> E is Young’s modulus.The wave equation can be generalised slightly. For example, in the case of thevibrating string, there could also be an external upward vertical <strong>for</strong>ce f(x, t) perunit length acting on the string at time t. The transverse vibrations would thensatisfy the equationT ∂2 u∂x 2 + f(x, t) u=ρ∂2 ∂t 2 ,which is clearly of the <strong>for</strong>m ‘upward <strong>for</strong>ce per unit length = mass per unit length× upward acceleration’.Similar examples, but involving two or three spatial dimensions rather than one,are provided by the equation governing the transverse vibrations of a stretchedmembrane subject to an external vertical <strong>for</strong>ce density f(x, y, t),( ∂ 2 )uT∂x 2 + ∂2 u∂y 2 + f(x, y, t) =ρ(x, y) ∂2 u∂t 2 ,where ρ is the mass per unit area of the membrane <strong>and</strong> T is the tension.677


PDES: GENERAL AND PARTICULAR SOLUTIONSThe diffusion equation20.1.2 The diffusion equationκ∇ 2 u = ∂u(20.2)∂tdescribes the temperature u in a region containing no heat sources or sinks; italso applies to the diffusion of a chemical that has a concentration u(r,t). Theconstant κ is called the diffusivity. The equation is clearly second order in thethree spatial variables, but first order in time.◮Derive the equation satisfied by the temperature u(r,t) at time t <strong>for</strong> a material of uni<strong>for</strong>mthermal conductivity k, specific heat capacity s <strong>and</strong> density ρ. Express the equation inCartesian coordinates.Let us consider an arbitrary volume V lying within the solid <strong>and</strong> bounded by a surface S(this may coincide with the surface of the solid if so desired). At any point in the solidthe rate of heat flow per unit area in any given direction ˆr is proportional to minus thecomponent of the temperature gradient in that direction <strong>and</strong> so is given by (−k∇u) · ˆr. Thetotal flux of heat out of the volume V per unit time is given by− dQdt∫∫S= (−k∇u) · ˆn dS∫∫∫= ∇ · (−k∇u) dV , (20.3)Vwhere Q is the total heat energy in V at time t <strong>and</strong> ˆn is the outward-pointing unit normalto S; note that we have used the divergence theorem to convert the surface integral intoa volume integral.We can also express Q as a volume integral over V ,∫∫∫Q = sρu dV ,<strong>and</strong> its rate of change is then given byVdQdt = ∫∫∫Vsρ ∂u dV , (20.4)∂twhere we have taken the derivative with respect to time inside the integral (see section 5.12).Comparing (20.3) <strong>and</strong> (20.4), <strong>and</strong> remembering that the volume V is arbitrary, we obtainthe three-dimensional diffusion equationκ∇ 2 u = ∂u∂t ,where the diffusion coefficient κ = k/(sρ). To express this equation in Cartesian coordinates,we simply write ∇ 2 in terms of x, y <strong>and</strong> z to obtainκ( )∂ 2 u∂x + ∂2 u2 ∂y + ∂2 u= ∂u2 ∂z 2 ∂t . ◭The diffusion equation just derived can be generalised tok∇ 2 u + f(r,t)=sρ ∂u∂t .678


20.1 IMPORTANT PARTIAL DIFFERENTIAL EQUATIONSThe second term, f(r,t), represents a varying density of heat sources throughoutthe material but is often not required in physical applications. In the most generalcase, k, s <strong>and</strong> ρ may depend on position r, in which case the first term becomes∇ · (k∇u). However, in the simplest application the heat flow is one-dimensionalwith no heat sources, <strong>and</strong> the equation becomes (in Cartesian coordinates)∂ 2 u∂x 2 = sρ ∂uk ∂t .Laplace’s equation,20.1.3 Laplace’s equation∇ 2 u =0, (20.5)may be obtained by setting ∂u/∂t = 0 in the diffusion equation (20.2), <strong>and</strong>describes (<strong>for</strong> example) the steady-state temperature distribution in a solid inwhich there are no heat sources – i.e. the temperature distribution after a longtime has elapsed.Laplace’s equation also describes the gravitational potential in a region containingno matter or the electrostatic potential in a charge-free region. Further, itapplies to the flow of an incompressible fluid with no sources, sinks or vortices;in this case u is the velocity potential, from which the velocity is given by v = ∇u.Poisson’s equation,20.1.4 Poisson’s equation∇ 2 u = ρ(r), (20.6)describes the same physical situations as Laplace’s equation, but in regionscontaining matter, charges or sources of heat or fluid. The function ρ(r) iscalled the source density <strong>and</strong> in physical applications usually contains somemultiplicative physical constants. For example, if u is the electrostatic potentialin some region of space, in which case ρ is the density of electric charge, then∇ 2 u = −ρ(r)/ɛ 0 ,whereɛ 0 is the permittivity of free space. Alternatively, u mightrepresent the gravitational potential in some region where the matter density isgiven by ρ; then∇ 2 u =4πGρ(r), where G is the gravitational constant.The Schrödinger equation20.1.5 Schr-odinger’s equation− 22m ∇2 u + V (r)u = i ∂u∂t , (20.7)679


PDES: GENERAL AND PARTICULAR SOLUTIONSdescribes the quantum mechanical wavefunction u(r,t) of a non-relativistic particleof mass m; is Planck’s constant divided by 2π. Like the diffusion equation it issecond order in the three spatial variables <strong>and</strong> first order in time.20.2 General <strong>for</strong>m of solutionBe<strong>for</strong>e turning to the methods by which we may hope to solve PDEs such asthose listed in the previous section, it is instructive, as <strong>for</strong> ODEs in chapter 14, tostudy how PDEs may be <strong>for</strong>med from a set of possible solutions. Such a studycan provide an indication of how equations obtained not from possible solutionsbut from physical arguments might be solved.For definiteness let us suppose we have a set of functions involving twoindependent variables x <strong>and</strong> y. Without further specification this is of course avery wide set of functions, <strong>and</strong> we could not expect to find a useful equation thatthey all satisfy. However, let us consider a type of function u i (x, y) inwhichx <strong>and</strong>y appear in a particular way, such that u i can be written as a function (howevercomplicated) of a single variable p, itself a simple function of x <strong>and</strong> y.Let us illustrate this by considering the three functionsu 1 (x, y) =x 4 +4(x 2 y + y 2 +1),u 2 (x, y) =sinx 2 cos 2y +cosx 2 sin 2y,u 3 (x, y) = x2 +2y +23x 2 +6y +5 .These are all fairly complicated functions of x <strong>and</strong> y <strong>and</strong> a single differentialequation of which each one is a solution is not obvious. However, if we observethat in fact each can be expressed as a function of the variable p = x 2 +2y alone(withnootherx or y involved) then a great simplification takes place. Writtenin terms of p the above equations becomeu 1 (x, y) =(x 2 +2y) 2 +4=p 2 +4=f 1 (p),u 2 (x, y) =sin(x 2 +2y) =sinp = f 2 (p),u 3 (x, y) = (x2 +2y)+23(x 2 +2y)+5 = p +23p +5 = f 3(p).Let us now <strong>for</strong>m, <strong>for</strong> each u i , the partial derivatives ∂u i /∂x <strong>and</strong> ∂u i /∂y. Ineachcase these are (writing both the <strong>for</strong>m <strong>for</strong> general p <strong>and</strong> the one appropriate toour particular case, p = x 2 +2y)∂u i∂x = df i(p) ∂pdp ∂x =2xf′ i,∂u i∂y = df i(p) ∂pdp ∂y =2f′ i,<strong>for</strong> i = 1, 2, 3. All reference to the <strong>for</strong>m of f i can be eliminated from these680


20.3 GENERAL AND PARTICULAR SOLUTIONSequations by cross-multiplication, obtaining∂p ∂u i∂yor, <strong>for</strong> our specific <strong>for</strong>m, p = x 2 +2y,∂x = ∂p∂x∂u i∂y ,∂u i∂x = x∂u i∂y . (20.8)It is thus apparent that not only are the three functions u 1 , u 2 u 3 solutions of thePDE (20.8) but so also is any arbitrary function f(p) of which the argument p hasthe <strong>for</strong>m x 2 +2y.20.3 General <strong>and</strong> particular solutionsIn the last section we found that the first-order PDE (20.8) has as a solution anyfunction of the variable x 2 +2y. This points the way <strong>for</strong> the solution of PDEsof other orders, as follows. It is not generally true that an nth-order PDE canalways be considered as resulting from the elimination of n arbitrary functionsfrom its solution (as opposed to the elimination of n arbitrary constants <strong>for</strong> annth-order ODE, see section 14.1). However, given specific PDEs we can try tosolve them by seeking combinations of variables in terms of which the solutionsmay be expressed as arbitrary functions. Where this is possible we may expect ncombinations to be involved in the solution.Naturally, the exact functional <strong>for</strong>m of the solution <strong>for</strong> any particular situationmust be determined by some set of boundary conditions. For instance, if the PDEcontains two independent variables x <strong>and</strong> y then <strong>for</strong> complete determination ofits solution the boundary conditions will take a <strong>for</strong>m equivalent to specifyingu(x, y) along a suitable continuum of points in the xy-plane (usually along a line).We now discuss the general <strong>and</strong> particular solutions of first- <strong>and</strong> secondorderPDEs. In order to simplify the algebra, we will restrict our discussionto equations containing just two independent variables x <strong>and</strong> y. Nevertheless,the method presented below may be extended to equations containing severalindependent variables.20.3.1 First-order equationsAlthough most of the PDEs encountered in physical contexts are second order(i.e. they contain ∂ 2 u/∂x 2 or ∂ 2 u/∂x∂y, etc.), we now discuss first-order equationsto illustrate the general considerations involved in the <strong>for</strong>m of the solution <strong>and</strong>in satisfying any boundary conditions on the solution.The most general first-order linear PDE (containing two independent variables)681


PDES: GENERAL AND PARTICULAR SOLUTIONSis of the <strong>for</strong>mA(x, y) ∂u ∂u+ B(x, y) + C(x, y)u = R(x, y), (20.9)∂x ∂ywhere A(x, y), B(x, y), C(x, y) <strong>and</strong>R(x, y) are given functions. Clearly, if eitherA(x, y) orB(x, y) is zero then the PDE may be solved straight<strong>for</strong>wardly as afirst-order linear ODE (as discussed in chapter 14), the only modification beingthat the arbitrary constant of integration becomes an arbitrary function of x or yrespectively.◮Find the general solution u(x, y) ofx ∂u∂x +3u = x2 .Dividing through by x we obtain∂u∂x + 3ux = x,which is a linear equation with integrating factor (see subsection 14.2.4)(∫ ) 3expx dx =exp(3lnx) =x 3 .Multiplying through by this factor we find∂∂x (x3 u)=x 4 ,which, on integrating with respect to x, givesx 3 u = x55 + f(y),where f(y) isanarbitrary function of y. Finally, dividing through by x 3 , we obtain thesolutionu(x, y) = x25 + f(y)x . ◭ 3When the PDE contains partial derivatives with respect to both independentvariables then, of course, we cannot employ the above procedure but must seekan alternative method. Let us <strong>for</strong> the moment restrict our attention to the specialcase in which C(x, y) =R(x, y) = 0 <strong>and</strong>, following the discussion of the previoussection, look <strong>for</strong> solutions of the <strong>for</strong>m u(x, y) =f(p) wherep is some, at presentunknown, combination of x <strong>and</strong> y. We then have∂u∂x = df(p) ∂pdp ∂x ,∂u∂y = df(p) ∂pdp ∂y ,682


20.3 GENERAL AND PARTICULAR SOLUTIONSwhich, when substituted into the PDE (20.9), give[A(x, y) ∂p]∂p df(p)+ B(x, y) =0.∂x ∂y dpThis removes all reference to the actual <strong>for</strong>m of the function f(p) since <strong>for</strong>non-trivial p we must haveA(x, y) ∂p ∂p+ B(x, y) =0. (20.10)∂x ∂yLet us now consider the necessary condition <strong>for</strong> f(p) to remain constant as x<strong>and</strong> y vary; this is that p itself remains constant. Thus <strong>for</strong> f to remain constantimplies that x <strong>and</strong> y must vary in such a way thatdp = ∂p ∂pdx + dy =0. (20.11)∂x ∂yThe <strong>for</strong>ms of (20.10) <strong>and</strong> (20.11) are very alike <strong>and</strong> become the same if werequire thatdxA(x, y) = dyB(x, y) . (20.12)By integrating this expression the <strong>for</strong>m of p can be found.◮Forx ∂u ∂u− 2y =0, (20.13)∂x ∂yfind (i) the solution that takes the value 2y +1 on the line x =1, <strong>and</strong> (ii) a solution thathas the value 4 at the point (1, 1).If we seek a solution of the <strong>for</strong>m u(x, y) =f(p), we deduce from (20.12) that u(x, y) willbe constant along lines of (x, y) thatsatisfydxx = dy−2y ,which on integrating gives x = cy −1/2 . Identifying the constant of integration c with p 1/2(to avoid fractional powers), we conclude that p = x 2 y. Thus the general solution of thePDE (20.13) isu(x, y) =f(x 2 y),where f is an arbitrary function.We must now find the particular solutions that obey each of the imposed boundaryconditions. For boundary condition (i) a little thought shows that the particular solutionrequired isu(x, y) =2(x 2 y)+1=2x 2 y +1. (20.14)For boundary condition (ii) some obviously acceptable solutions areu(x, y) =x 2 y +3,u(x, y) =4x 2 y,u(x, y) =4.683


PDES: GENERAL AND PARTICULAR SOLUTIONSEach is a valid solution (the freedom of choice of <strong>for</strong>m arises from the fact that uis specified at only one point (1, 1), <strong>and</strong> not along a continuum (say), as in boundarycondition (i)). All three are particular examples of the general solution, which may bewritten, <strong>for</strong> example, asu(x, y) =x 2 y +3+g(x 2 y),where g = g(x 2 y)=g(p) is an arbitrary function subject only to g(1) = 0. For thisexample, the <strong>for</strong>ms of g corresponding to the particular solutions listed above are g(p) =0,g(p) =3p − 3, g(p) =1− p. ◭As mentioned above, in order to find a solution of the <strong>for</strong>m u(x, y) =f(p) werequire that the original PDE contains no term in u, but only terms containingits partial derivatives. If a term in u is present, so that C(x, y) ≠ 0 in (20.9),then the procedure needs some modification, since we cannot simply divide outthe dependence on f(p) to obtain (20.10). In such cases we look instead <strong>for</strong>a solution of the <strong>for</strong>m u(x, y) =h(x, y)f(p). We illustrate this method in thefollowing example.◮Find the general solution ofx ∂u∂x +2∂u − 2u =0. (20.15)∂yWe seek a solution of the <strong>for</strong>m u(x, y) =h(x, y)f(p), with the consequence that∂u∂x = ∂h df(p) ∂pf(p)+h∂x dp ∂x ,∂u∂y = ∂h df(p) ∂pf(p)+h∂y dp ∂y .Substituting these expressions into the PDE (20.15) <strong>and</strong> rearranging, we obtain(x ∂h) (∂x +2∂h ∂y − 2h f(p)+ x ∂p )∂x +2∂p h df(p) =0.∂y dpThe first factor in parentheses is just the original PDE with u replaced by h. There<strong>for</strong>e, ifh is any solution of the PDE, however simple, this term will vanish, to leave(x ∂p )∂x +2∂p h df(p) =0,∂y dpfrom which, as in the previous case, we obtainx ∂p∂x +2∂p ∂y =0.From (20.11) <strong>and</strong> (20.12) we see that u(x, y) will be constant along lines of (x, y) thatsatisfydxx = dy 2 ,which integrates to give x = c exp(y/2). Identifying the constant of integration c with p wefind p = x exp(−y/2). Thus the general solution of (20.15) isu(x, y) =h(x, y)f(x exp(− 1 y)), 2where f(p) is any arbitrary function of p <strong>and</strong> h(x, y) is any solution of (20.15).684


20.3 GENERAL AND PARTICULAR SOLUTIONSIf we take, <strong>for</strong> example, h(x, y) =expy, which clearly satisfies (20.15), then the generalsolution isu(x, y) =(expy)f(x exp(− 1 y)). 2Alternatively, h(x, y) =x 2 also satisfies (20.15) <strong>and</strong> so the general solution to the equationcan also be writtenu(x, y) =x 2 g(x exp(− 1 y)), 2where g is an arbitrary function of p; clearly g(p) =f(p)/p 2 . ◭20.3.2 Inhomogeneous equations <strong>and</strong> problemsLet us discuss in a more general <strong>for</strong>m the particular solutions of (20.13) foundin the second example of the previous subsection. It is clear that, so far as thisequation is concerned, if u(x, y) is a solution then so is any multiple of u(x, y) orany linear sum of separate solutions u 1 (x, y)+u 2 (x, y). However, when it comesto fitting the boundary conditions this is not so.For example, although u(x, y) in (20.14) satisfies the PDE <strong>and</strong> the boundarycondition u(1,y)=2y + 1, the function u 1 (x, y) =4u(x, y) =8xy + 4, whilstsatisfying the PDE, takes the value 8y+4 on the line x = 1 <strong>and</strong> so does not satisfythe required boundary condition. Likewise the function u 2 (x, y) =u(x, y)+f 1 (x 2 y),<strong>for</strong> arbitrary f 1 , satisfies (20.13) but takes the value u 2 (1,y)=2y +1+f 1 (y) onthe line x = 1, <strong>and</strong> so is not of the required <strong>for</strong>m unless f 1 is identically zero.Thus we see that when treating the superposition of solutions of PDEs twoconsiderations arise, one concerning the equation itself <strong>and</strong> the other connectedto the boundary conditions. The equation is said to be homogeneous if the factthat u(x, y) is a solution implies that λu(x, y), <strong>for</strong> any constant λ, is also a solution.However, the problem is said to be homogeneous if, in addition, the boundaryconditions are such that if they are satisfied by u(x, y) then they are also satisfiedby λu(x, y). The last requirement itself is referred to as that of homogeneousboundary conditions.For example, the PDE (20.13) is homogeneous but the general first-orderequation (20.9) would not be homogeneous unless R(x, y) = 0. Furthermore,the boundary condition (i) imposed on the solution of (20.13) in the previoussubsection is not homogeneous though, in this case, the boundary conditionu(x, y) = 0on the line y =4x −2would be, since u(x, y) =λ(x 2 y − 4) satisfies this condition <strong>for</strong> any λ <strong>and</strong>, being afunction of x 2 y, satisfies (20.13).The reason <strong>for</strong> discussing the homogeneity of PDEs <strong>and</strong> their boundary conditionsis that in linear PDEs there is a close parallel to the complementary-function<strong>and</strong> particular-integral property of ODEs. The general solution of an inhomogeneousproblem can be written as the sum of any particular solution of theproblem <strong>and</strong> the general solution of the corresponding homogeneous problem (as685


PDES: GENERAL AND PARTICULAR SOLUTIONS<strong>for</strong> ODEs, we require that the particular solution is not already contained in thegeneral solution of the homogeneous problem). Thus, <strong>for</strong> example, the generalsolution of∂u∂x − x ∂u + au = f(x, y), (20.16)∂ysubject to, say, the boundary condition u(0,y)=g(y), is given byu(x, y) =v(x, y)+w(x, y),where v(x, y) is any solution (however simple) of (20.16) such that v(0,y)=g(y)<strong>and</strong> w(x, y) is the general solution of∂w∂x − x∂w + aw =0, (20.17)∂ywith w(0,y) = 0. If the boundary conditions are sufficiently specified then the onlypossible solution of (20.17) will be w(x, y) ≡ 0<strong>and</strong>v(x, y) will be the completesolution by itself.Alternatively, we may begin by finding the general solution of the inhomogeneousequation (20.16) without regard <strong>for</strong> any boundary conditions; it is just thesum of the general solution to the homogeneous equation <strong>and</strong> a particular integralof (20.16), both without reference to the boundary conditions. The boundaryconditions can then be used to find the appropriate particular solution from thegeneral solution.We will not discuss at length general methods of obtaining particular integralsof PDEs but merely note that some of those methods available <strong>for</strong> ordinarydifferential equations can be suitably extended. §◮Find the general solution ofy ∂u∂x − x ∂u =3x. (20.18)∂yHence find the most general particular solution (i) which satisfies u(x, 0) = x 2 <strong>and</strong> (ii) whichhas the value u(x, y) =2at the point (1, 0).This equation is inhomogeneous, <strong>and</strong> so let us first find the general solution of (20.18)without regard <strong>for</strong> any boundary conditions. We begin by looking <strong>for</strong> the solution of thecorresponding homogeneous equation ((20.18) but with the RHS equal to zero) of the<strong>for</strong>m u(x, y) =f(p). Following the same procedure as that used in the solution of (20.13)we find that u(x, y) will be constant along lines of (x, y) thatsatisfydxy = dy−x⇒x22 + y22 = c.Identifying the constant of integration c with p/2, we find that the general solution of the§ See <strong>for</strong> example H. T. H. Piaggio, An Elementary Treatise on Differential Equations <strong>and</strong> theirApplications (London: G. Bell <strong>and</strong> Sons, Ltd, 1954), pp. 175 ff.686


20.3 GENERAL AND PARTICULAR SOLUTIONShomogeneous equation is u(x, y) =f(x 2 + y 2 ) <strong>for</strong> arbitrary function f. Now by inspectiona particular integral of (20.18) is u(x, y) =−3y, <strong>and</strong> so the general solution to (20.18) isu(x, y) =f(x 2 + y 2 ) − 3y.Boundary condition (i) requires u(x, 0) = f(x 2 )=x 2 ,i.e.f(z) =z, <strong>and</strong> so the particularsolution in this case isu(x, y) =x 2 + y 2 − 3y.Similarly, boundary condition (ii) requires u(1, 0) = f(1) = 2. One possibility is f(z) =2z,<strong>and</strong> if we make this choice, then one way of writing the most general particular solutionisu(x, y) =2x 2 +2y 2 − 3y + g(x 2 + y 2 ),where g is any arbitrary function <strong>for</strong> which g(1) = 0. Alternatively, a simpler choice wouldbe f(z) = 2, leading tou(x, y) =2− 3y + g(x 2 + y 2 ). ◭Although we have discussed the solution of inhomogeneous problems only<strong>for</strong> first-order equations, the general considerations hold true <strong>for</strong> linear PDEs ofhigher order.20.3.3 Second-order equationsAs noted in section 20.1, second-order linear PDEs are of great importance indescribing the behaviour of many physical systems. As in our discussion of firstorderequations, <strong>for</strong> the moment we shall restrict our discussion to equations withjust two independent variables; extensions to a greater number of independentvariables are straight<strong>for</strong>ward.The most general second-order linear PDE (containing two independent variables)has the <strong>for</strong>mA ∂2 u∂x 2 + B ∂2 u∂x∂y + C ∂2 u∂y 2 + D ∂u∂x + E ∂u + Fu = R(x, y), (20.19)∂ywhere A,B,...,F <strong>and</strong> R(x, y) are given functions of x <strong>and</strong> y. Because of the natureof the solutions to such equations, they are usually divided into three classes, adivision of which we will make further use in subsection 20.6.2. The equation(20.19) is called hyperbolic if B 2 > 4AC, parabolic if B 2 =4AC <strong>and</strong> elliptic ifB 2 < 4AC. Clearly, if A, B <strong>and</strong> C are functions of x <strong>and</strong> y (rather than justconstants) then the equation might be of different types in different parts of thexy-plane.Equation (20.19) obviously represents a very large class of PDEs, <strong>and</strong> it isusually impossible to find closed-<strong>for</strong>m solutions to most of these equations.There<strong>for</strong>e, <strong>for</strong> the moment we shall consider only homogeneous equations, withR(x, y) = 0, <strong>and</strong> make the further (greatly simplifying) restriction that, throughoutthe remainder of this section, A,B,...,F are not functions of x <strong>and</strong> y but merelyconstants.687


PDES: GENERAL AND PARTICULAR SOLUTIONSWe now tackle the problem of solving some types of second-order PDE withconstant coefficients by seeking solutions that are arbitrary functions of particularcombinations of independent variables, just as we did <strong>for</strong> first-order equations.Following the discussion of the previous section, we can hope to find suchsolutions only if all the terms of the equation involve the same total numberof differentiations, i.e. all terms are of the same order, although the numberof differentiations with respect to the individual independent variables may bedifferent. This means that in (20.19) we require the constants D, E <strong>and</strong> F to beidentically zero (we have, of course, already assumed that R(x, y) iszero),sothatwe are now considering only equations of the <strong>for</strong>mA ∂2 u∂x 2 + B ∂2 u∂x∂y + C ∂2 u=0, (20.20)∂y2 where A, B <strong>and</strong> C are constants. We note that both the one-dimensional waveequation,∂ 2 u∂x 2 − 1 ∂ 2 uc 2 ∂t 2 =0,<strong>and</strong> the two-dimensional Laplace equation,∂ 2 u∂x 2 + ∂2 u∂y 2 =0,are of this <strong>for</strong>m, but that the diffusion equation,κ ∂2 u∂x 2 − ∂u∂t =0,is not, since it contains a first-order derivative.Since all the terms in (20.20) involve two differentiations, by assuming a solutionof the <strong>for</strong>m u(x, y) =f(p), where p is some unknown function of x <strong>and</strong> y (or t),we may be able to obtain a common factor d 2 f(p)/dp 2 as the only appearance off on the LHS. Then, because of the zero RHS, all reference to the <strong>for</strong>m of f canbe cancelled out.We can gain some guidance on suitable <strong>for</strong>ms <strong>for</strong> the combination p = p(x, y)by considering ∂u/∂x when u is given by u(x, y) =f(p), <strong>for</strong> then∂u∂x = df(p) ∂pdp ∂x .Clearly differentiation of this equation with respect to x (or y) will not lead to asingle term on the RHS, containing f only as d 2 f(p)/dp 2 , unless the factor ∂p/∂xis a constant so that ∂ 2 p/∂x 2 <strong>and</strong> ∂ 2 p/∂x∂y are necessarily zero. This shows thatp must be a linear function of x. In an exactly similar way p must also be a linearfunction of y, i.e.p = ax + by.If we assume a solution of (20.20) of the <strong>for</strong>m u(x, y) =f(ax+by), <strong>and</strong> evaluate688


20.3 GENERAL AND PARTICULAR SOLUTIONSthe terms ready <strong>for</strong> substitution into (20.20), we obtain∂ 2 u∂x 2 = a2 d2 f(p)dp 2 ,which on substitution give∂u∂x = adf(p) dp ,∂u∂y = bdf(p) dp ,∂ 2 u∂x∂y = f(p)abd2 dp 2 ,∂ 2 u∂y 2 = b2 d2 f(p)dp 2 ,(Aa 2 + Bab + Cb 2) d 2 f(p)dp 2 =0. (20.21)This is the <strong>for</strong>m we have been seeking, since now a solution independent ofthe <strong>for</strong>m of f can be obtained if we require that a <strong>and</strong> b satisfyAa 2 + Bab + Cb 2 =0.From this quadratic, two values <strong>for</strong> the ratio of the two constants a <strong>and</strong> b areobtained,b/a =[−B ± (B 2 − 4AC) 1/2 ]/2C.If we denote these two ratios by λ 1 <strong>and</strong> λ 2 then any functions of the two variablesp 1 = x + λ 1 y,p 2 = x + λ 2 ywill be solutions of the original equation (20.20). The omission of the constantfactor a from p 1 <strong>and</strong> p 2 is of no consequence since this can always be absorbedinto the particular <strong>for</strong>m of any chosen function; only the relative weighting of x<strong>and</strong> y in p is important.Since p 1 <strong>and</strong> p 2 are in general different, we can thus write the general solutionof (20.20) asu(x, y) =f(x + λ 1 y)+g(x + λ 2 y), (20.22)where f <strong>and</strong> g are arbitrary functions.Finally, we note that the alternative solution d 2 f(p)/dp 2 = 0 to (20.21) leadsonly to the trivial solution u(x, y) =kx + ly + m, <strong>for</strong> which all second derivativesare individually zero.◮ Find the general solution of the one-dimensional wave equation∂ 2 u∂x − 1 ∂ 2 u2 c 2 ∂t =0. 2This equation is (20.20) with A =1,B =0<strong>and</strong>C = −1/c 2 , <strong>and</strong> so the values of λ 1 <strong>and</strong> λ 2are the solutions of1 − λ2c =0, 2namely λ 1 = −c <strong>and</strong> λ 2 = c. This means that arbitrary functions of the quantitiesp 1 = x − ct, p 2 = x + ct689


PDES: GENERAL AND PARTICULAR SOLUTIONSwill be satisfactory solutions of the equation <strong>and</strong> that the general solution will beu(x, t) =f(x − ct)+g(x + ct), (20.23)where f <strong>and</strong> g are arbitrary functions. This solution is discussed further in section 20.4. ◭The method used to obtain the general solution of the wave equation may alsobe applied straight<strong>for</strong>wardly to Laplace’s equation.◮ Find the general solution of the two-dimensional Laplace equation∂ 2 u∂x + ∂2 u=0. (20.24)2 ∂y2 Following the established procedure, we look <strong>for</strong> a solution that is a function f(p) ofp = x + λy, where from (20.24) λ satisfies1+λ 2 =0.This requires that λ = ±i, <strong>and</strong> satisfactory variables p are p = x ± iy. The general solutionrequired is there<strong>for</strong>e, in terms of arbitrary functions f <strong>and</strong> g,u(x, y) =f(x + iy)+g(x − iy). ◭It will be apparent from the last two examples that the nature of the appropriatelinear combination of x <strong>and</strong> y depends upon whether B 2 > 4AC or B 2 < 4AC.This is exactly the same criterion as determines whether the PDE is hyperbolicor elliptic. Hence as a general result, hyperbolic <strong>and</strong> elliptic equations of the<strong>for</strong>m (20.20), given the restriction that the constants A, B <strong>and</strong> C are real, have assolutions functions whose arguments have the <strong>for</strong>m x+αy <strong>and</strong> x+iβy respectively,where α <strong>and</strong> β themselves are real.The one case not covered by this result is that in which B 2 =4AC, i.e.aparabolic equation. In this case λ 1 <strong>and</strong> λ 2 are not different <strong>and</strong> only one suitablecombination of x <strong>and</strong> y results, namelyu(x, y) =f(x − (B/2C)y).To find the second part of the general solution we try, in analogy with thecorresponding situation <strong>for</strong> ordinary differential equations, a solution of the <strong>for</strong>mu(x, y) =h(x, y)g(x − (B/2C)y).Substituting this into (20.20) <strong>and</strong> using A = B 2 /4C results in()A ∂2 h∂x 2 + B ∂2 h∂x∂y + C ∂2 h∂y 2 g =0.There<strong>for</strong>e we require h(x, y) to be any solution of the original PDE. There areseveral simple solutions of this equation, but as only one is required we take thesimplest non-trivial one, h(x, y) =x, to give the general solution of the parabolicequationu(x, y) =f(x − (B/2C)y)+xg(x − (B/2C)y). (20.25)690


20.3 GENERAL AND PARTICULAR SOLUTIONSWe could, of course, have taken h(x, y) =y, but this only leads to a solution thatis already contained in (20.25).◮Solve∂ 2 u∂x +2 ∂2 u2 ∂x∂y + ∂2 u∂y =0, 2subject to the boundary conditions u(0,y)=0<strong>and</strong> u(x, 1) = x 2 .From our general result, functions of p = x + λy will be solutions provided1+2λ + λ 2 =0,i.e. λ = −1 <strong>and</strong> the equation is parabolic. The general solution is there<strong>for</strong>eu(x, y) =f(x − y)+xg(x − y).The boundary condition u(0,y) = 0 implies f(p) ≡ 0, whilst u(x, 1) = x 2 yieldsxg(x − 1) = x 2 ,which gives g(p) =p + 1, There<strong>for</strong>e the particular solution required isu(x, y) =x(p +1)=x(x − y +1). ◭To rein<strong>for</strong>ce the material discussed above we will now give alternative derivationsof the general solutions (20.22) <strong>and</strong> (20.25) by expressing the original PDEin terms of new variables be<strong>for</strong>e solving it. The actual solution will then becomealmost trivial; but, of course, it will be recognised that suitable new variablescould hardly have been guessed if it were not <strong>for</strong> the work already done. Thisdoes not detract from the validity of the derivation to be described, only fromthe likelihood that it would be discovered by inspection.We start again with (20.20) <strong>and</strong> change to new variablesζ = x + λ 1 y, η = x + λ 2 y.With this change of variables, we have from the chain rule thatUsing these <strong>and</strong> the fact that∂∂x = ∂∂ζ + ∂∂η ,∂∂y = λ ∂1∂ζ + λ ∂2∂η .equation (20.20) becomesA + Bλ i + Cλ 2 i =0 <strong>for</strong>i =1, 2,[2A + B(λ 1 + λ 2 )+2Cλ 1 λ 2 ] ∂2 u∂ζ∂η =0.691


PDES: GENERAL AND PARTICULAR SOLUTIONSThen, providing the factor in brackets does not vanish, <strong>for</strong> which the requiredcondition is easily shown to be B 2 ≠4AC, weobtainwhich has the successive integrals∂ 2 u∂ζ∂η =0,∂u∂η = F(η),u(ζ,η)=f(η)+g(ζ).This solution is just the same as (20.22),u(x, y) =f(x + λ 2 y)+g(x + λ 1 y).If the equation is parabolic (i.e. B 2 =4AC), we instead use the new variablesζ = x + λy, η = x,<strong>and</strong> recalling that λ = −(B/2C) we can reduce (20.20) toA ∂2 u∂η 2 =0.Two straight<strong>for</strong>ward integrations give as the general solutionu(ζ,η)=ηg(ζ)+f(ζ),whichintermsofx <strong>and</strong> y has exactly the <strong>for</strong>m of (20.25),u(x, y) =xg(x + λy)+f(x + λy).Finally, as hinted at in subsection 20.3.2 with reference to first-order linearPDEs, some of the methods used to find particular integrals of linear ODEscan be suitably modified to find particular integrals of PDEs of higher order. Insimple cases, however, an appropriate solution may often be found by inspection.◮Find the general solution of∂ 2 u∂x + ∂2 u=6(x + y).2 ∂y2 Following our previous methods <strong>and</strong> results, the complementary function isu(x, y) =f(x + iy)+g(x − iy),<strong>and</strong> only a particular integral remains to be found. By inspection a particular integral ofthe equation is u(x, y) =x 3 + y 3 , <strong>and</strong> so the general solution can be writtenu(x, y) =f(x + iy)+g(x − iy)+x 3 + y 3 . ◭692


20.4 THE WAVE EQUATION20.4 The wave equationWe have already found that the general solution of the one-dimensional waveequation isu(x, t) =f(x − ct)+g(x + ct), (20.26)where f <strong>and</strong> g are arbitrary functions. However, the equation is of such generalimportance that further discussion will not be out of place.Let us imagine that u(x, t) =f(x − ct) represents the displacement of a string attime t <strong>and</strong> position x. It is clear that all positions x <strong>and</strong> times t <strong>for</strong> which x − ct =constant will have the same instantaneous displacement. But x − ct = constantis exactly the relation between the time <strong>and</strong> position of an observer travellingwith speed c along the positive x-direction. Consequently this moving observersees a constant displacement of the string, whereas to a stationary observer, theinitial profile u(x, 0) moves with speed c along the x-axis as if it were a rigidsystem. Thus f(x − ct) represents a wave <strong>for</strong>m of constant shape travelling alongthe positive x-axis with speed c, the actual <strong>for</strong>m of the wave depending uponthe function f. Similarly, the term g(x + ct) is a constant wave <strong>for</strong>m travellingwith speed c in the negative x-direction. The general solution (20.23) representsa superposition of these.If the functions f <strong>and</strong> g are the same then the complete solution (20.23)represents identical progressive waves going in opposite directions. This mayresult in a wave pattern whose profile does not progress, described as a st<strong>and</strong>ingwave. As a simple example, suppose both f(p) <strong>and</strong>g(p) have the <strong>for</strong>m §Then (20.23) can be written asf(p) =g(p) =A cos(kp + ɛ).u(x, t) =A[cos(kx − kct + ɛ)+cos(kx + kct + ɛ)]=2A cos(kct)cos(kx + ɛ).The important thing to notice is that the shape of the wave pattern, given by thefactor in x, is the same at all times but that its amplitude 2A cos(kct) dependsupon time. At some points x that satisfycos(kx + ɛ) =0there is no displacement at any time; such points are called nodes.So far we have not imposed any boundary conditions on the solution (20.26).The problem of finding a solution to the wave equation that satisfies given boundaryconditions is normally treated using the method of separation of variables§ In the usual notation, k is the wave number (= 2π/wavelength) <strong>and</strong> kc = ω, the angular frequencyof the wave.693


PDES: GENERAL AND PARTICULAR SOLUTIONSdiscussed in the next chapter. Nevertheless, we now consider D’Alembert’s solutionu(x, t) of the wave equation subject to initial conditions (boundary conditions) inthe following general <strong>for</strong>m:∂u(x, 0)initial displacement, u(x, 0) = φ(x); initial velocity, = ψ(x).∂tThe functions φ(x) <strong>and</strong>ψ(x) are given <strong>and</strong> describe the displacement <strong>and</strong> velocityof each part of the string at the (arbitrary) time t =0.It is clear that what we need are the particular <strong>for</strong>ms of the functions f <strong>and</strong> gin (20.26) that lead to the required values at t = 0. This means thatφ(x) =u(x, 0) = f(x − 0) + g(x +0), (20.27)∂u(x, 0)ψ(x) = = −cf ′ (x − 0) + cg ′ (x +0), (20.28)∂twhere it should be noted that f ′ (x − 0) st<strong>and</strong>s <strong>for</strong> df(p)/dp evaluated, after thedifferentiation, at p = x − c × 0; likewise <strong>for</strong> g ′ (x +0).Looking on the above two left-h<strong>and</strong> sides as functions of p = x ± ct, buteverywhere evaluated at t = 0, we may integrate (20.28) between an arbitrary(<strong>and</strong> irrelevant) lower limit p 0 <strong>and</strong> an indefinite upper limit p to obtain∫1 pψ(q) dq + K = −f(p)+g(p),c p 0the constant of integration K depending on p 0 . Comparing this equation with(20.27), with x replaced by p, we can establish the <strong>for</strong>ms of the functions f <strong>and</strong>g asf(p) = φ(p) − 1 ∫ pψ(q) dq − K 2 2c p 02 , (20.29)g(p) = φ(p) + 1 ∫ pψ(q) dq + K 2 2c p 02 . (20.30)Adding (20.29) with p = x − ct to (20.30) with p = x + ct gives as the solution tothe original problem∫ x+ctu(x, t) = 1 2 [φ(x − ct)+φ(x + ct)] + 1 ψ(q) dq, (20.31)2c x−ctin which we notice that all dependence on p 0 has disappeared.Each of the terms in (20.31) has a fairly straight<strong>for</strong>ward physical interpretation.In each case the factor 1/2 represents the fact that only half a displacement profilethat starts at any particular point on the string travels towards any other positionx, the other half travelling away from it. The first term 1 2φ(x − ct) arises fromthe initial displacement at a distance ct to the left of x; this travels <strong>for</strong>wardarriving at x at time t. Similarly, the second contribution is due to the initialdisplacement at a distance ct to the right of x. The interpretation of the final694


20.5 THE DIFFUSION EQUATIONterm is a little less obvious. It can be viewed as representing the accumulatedtransverse displacement at position x due to the passage past x of all parts ofthe initial motion whose effects can reach x within a time t, both backward <strong>and</strong><strong>for</strong>ward travelling.The extension to the three-dimensional wave equation of solutions of the typewe have so far encountered presents no serious difficulty. In Cartesian coordinatesthe three-dimensional wave equation is∂ 2 u∂x 2 + ∂2 u∂y 2 + ∂2 u∂z 2 − 1 ∂ 2 uc 2 =0. (20.32)∂t2 In close analogy with the one-dimensional case we try solutions that are functionsof linear combinations of all four variables,p = lx + my + nz + µt.It is clear that a solution u(x, y, z, t) =f(p) will be acceptable provided that) (l 2 + m 2 + n 2 − µ2 d 2 f(p)c 2 dp 2 =0.Thus, as in the one-dimensional case, f can be arbitrary provided thatl 2 + m 2 + n 2 = µ 2 /c 2 .Using an obvious normalisation, we take µ = ±c <strong>and</strong> l, m, n as three numberssuch thatl 2 + m 2 + n 2 =1.In other words (l,m,n) are the Cartesian components of a unit vector ˆn thatpoints along the direction of propagation of the wave. The quantity p can bewritten in terms of vectors as the scalar expression p = ˆn · r ± ct, <strong>and</strong> the generalsolution of (20.32) is thenu(x, y, z, t) =u(r,t)=f(ˆn · r − ct)+g(ˆn · r + ct), (20.33)where ˆn is any unit vector. It would perhaps be more transparent to write ˆnexplicitly as one of the arguments of u.20.5 The diffusion equationOne important class of second-order PDEs, which we have not yet consideredin detail, is that in which the second derivative with respect to one variableappears, but only the first derivative with respect to another (usually time). Thisis exemplified by the one-dimensional diffusion equationκ ∂2 u(x, t)∂x 2695= ∂u∂t , (20.34)


PDES: GENERAL AND PARTICULAR SOLUTIONSin which κ is a constant with the dimensions length 2 × time −1 . The physicalconstants that go to make up κ in a particular case depend upon the nature ofthe process (e.g. solute diffusion, heat flow, etc.) <strong>and</strong> the material being described.With (20.34) we cannot hope to repeat successfully the method of subsection20.3.3, since now u(x, t) is differentiated a different number of times on the twosides of the equation; any attempted solution in the <strong>for</strong>m u(x, t) =f(p) withp = ax + bt will lead only to an equation in which the <strong>for</strong>m of f cannot becancelled out. Clearly we must try other methods.Solutions may be obtained by using the st<strong>and</strong>ard method of separation ofvariables discussed in the next chapter. Alternatively, a simple solution is alsogiven if both sides of (20.34), as it st<strong>and</strong>s, are separately set equal to a constantα (say), so that∂ 2 u∂x 2 = α κ ,These equations have the general solutions∂u∂t = α.u(x, t) = α2κ x2 + xg(t)+h(t) <strong>and</strong> u(x, t) =αt + m(x)respectively <strong>and</strong> may be made compatible with each other if g(t) is taken asconstant, g(t) =g (where g could be zero), h(t) =αt <strong>and</strong> m(x) =(α/2κ)x 2 + gx.An acceptable solution is thusu(x, t) = α2κ x2 + gx + αt + constant. (20.35)Let us now return to seeking solutions of equations by combining the independentvariables in particular ways. Having seen that a linear combination ofx <strong>and</strong> t will be of no value, we must search <strong>for</strong> other possible combinations. Ithas been noted already that κ has the dimensions length 2 × time −1 <strong>and</strong> so thecombination of variablesη = x2κtwill be dimensionless. Let us see if we can satisfy (20.34) with a solution of the<strong>for</strong>m u(x, t) =f(η). Evaluating the necessary derivatives we have∂u∂x = df(η) ∂ηdη ∂x = 2x df(η)κt dη ,∂ 2 u∂x 2 = 2 (df(η) 2xκt dη+ κt) 2d 2 f(η)dη 2 ,∂u∂t = − x2 df(η)κt 2 dη .Substituting these expressions into (20.34) we find that the new equation can be696


20.5 THE DIFFUSION EQUATIONwritten entirely in terms of η,4η d2 f(η)dη 2+(2+η) df(η)dηThis is a straight<strong>for</strong>ward ODE, which can be solved as follows. Writing f ′ (η) =df(η)/dη, etc., we havef ′′ (η)f ′ (η) = − 12η − 1 4⇒ ln[η 1/2 f ′ (η)] = − η 4 + c⇒ f ′ (η) = A ( −η)η exp 1/2 4∫ η ( −µ)⇒ f(η) =A µ −1/2 exp dµ.η 04If we now write this in terms of a slightly different variableζ = η1/22 = x2(κt) , 1/2then dζ = 1 4 η−1/2 dη, <strong>and</strong> the solution to (20.34) is given by∫ ζu(x, t) =f(η) =g(ζ) =B exp(−ν 2 ) dν. (20.36)ζ 0Here B is a constant <strong>and</strong> it should be noticed that x <strong>and</strong> t appear on the RHSonly in the indefinite upper limit ζ, <strong>and</strong> then only in the combination xt −1/2 .Ifζ 0 is chosen as zero then u(x, t) is, to within a constant factor, § the error functionerf[x/2(κt) 1/2 ], which is tabulated in many reference books. Only non-negativevalues of x <strong>and</strong> t are to be considered here, so that ζ ≥ ζ 0 .Let us try to determine what kind of (say) temperature distribution <strong>and</strong> flowthis represents. For definiteness we take ζ 0 = 0. Firstly, since u(x, t) in (20.36)depends only upon the product xt −1/2 , it is clear that all points x at times t suchthat xt −1/2 has the same value have the same temperature. Put another way, atany specific time t the region having a particular temperature has moved alongthe positive x-axis a distance proportional to the square root of t. Thisisatypicaldiffusion process.Notice that, on the one h<strong>and</strong>, at t = 0 the variable ζ →∞<strong>and</strong> u becomesquite independent of x (except perhaps at x = 0); the solution then represents auni<strong>for</strong>m spatial temperature distribution. On the other h<strong>and</strong>, at x = 0 we havethat u(x, t) is identically zero <strong>for</strong> all t.=0.§ Take B =2π −1/2 to give the usual error function normalised in such a way that erf(∞) =1.Seethe Appendix.697


PDES: GENERAL AND PARTICULAR SOLUTIONS◮An infrared laser delivers a pulse of (heat) energy E to a point P on a large insulatedsheet of thickness b, thermal conductivity k, specific heat s <strong>and</strong> density ρ. The sheet isinitially at a uni<strong>for</strong>m temperature. If u(r, t) is the excess temperature a time t later, at apoint that is a distance r (≫ b) from P , then show that a suitable expression <strong>for</strong> u isu(r, t) = α t exp (− r22βt), (20.37)where α <strong>and</strong> β are constants. (Note that we use r instead of ρ to denote the radial coordinatein plane polars so as to avoid confusion with the density.)Further, (i) show that β =2k/(sρ); (ii) demonstrate that the excess heat energy in thesheet is independent of t, <strong>and</strong> hence evaluate α; <strong>and</strong> (iii) prove that the total heat flow pastany circle of radius r is E.The equation to be solved is the heat diffusion equationk∇ 2 u(r,t)=sρ ∂u(r,t) .∂tSince we only require the solution <strong>for</strong> r ≫ b we can treat the problem as two-dimensionalwith obvious circular symmetry. Thus only the r-derivative term in the expression <strong>for</strong> ∇ 2 uis non-zero, giving(k ∂r ∂u )= sρ ∂ur ∂r ∂r ∂t , (20.38)where now u(r,t)=u(r, t).(i) Substituting the given expression (20.37) into (20.38) we obtain( ))2kα r2βt 2 2βt − 1 exp(− r2= sρα ( ) r22βt t 2 2βt − 1 exp(− r22βtfrom which we find that (20.37) is a solution, provided β =2k/(sρ).(ii) The excess heat in the system at any time t is∫ ∞∫ ∞( )rbρs u(r, t)2πr dr =2πbρsα00 t exp − r2dr2βt=2πbρsαβ.The excess heat is there<strong>for</strong>e independent of t <strong>and</strong> so must be equal to the total heatinput E, implying thatEα =2πbρsβ = E4πbk .(iii) The total heat flow past a circle of radius r is−2πrbk∫ ∞0∂u(r, t)dt = −2πrbk∂r[= E exp∫ ∞0(− r2E4πbkt),( ))−rexp(− r2dtβt 2βt∞= E <strong>for</strong> all r.2βt)]0As we would expect, all the heat energy E deposited by the laser will eventually flow pasta circle of any given radius r. ◭698


20.6 CHARACTERISTICS AND THE EXISTENCE OF SOLUTIONS20.6 Characteristics <strong>and</strong> the existence of solutionsSo far in this chapter we have discussed how to find general solutions to varioustypes of first- <strong>and</strong> second-order linear PDE. Moreover, given a set of boundaryconditions we have shown how to find the particular solution (or class of solutions)that satisfies them. For first-order equations, <strong>for</strong> example, we found that if thevalue of u(x, y) is specified along some curve in the xy-plane then the solution tothe PDE is in general unique, but that if u(x, y) is specified at only a single pointthen the solution is not unique: there exists a class of particular solutions all ofwhich satisfy the boundary condition. In this section <strong>and</strong> the next we make morerigorous the notion of the respective types of boundary condition that cause aPDE to have a unique solution, a class of solutions, or no solution at all.20.6.1 First-order equationsLet us consider the general first-order PDE (20.9) but now write it asA(x, y) ∂u ∂u+ B(x, y) = F(x, y, u). (20.39)∂x ∂ySuppose we wish to solve this PDE subject to the boundary condition thatu(x, y) =φ(s) is specified along some curve C in the xy-plane that is describedparametrically by the equations x = x(s) <strong>and</strong>y = y(s), where s is the arc lengthalong C. The variation of u along C is there<strong>for</strong>e given byduds = ∂u dx∂x ds + ∂u dy∂y ds = dφds . (20.40)We may then solve the two (inhomogeneous) simultaneous linear equations(20.39) <strong>and</strong> (20.40) <strong>for</strong> ∂u/∂x <strong>and</strong> ∂u/∂y, unless the determinant of the coefficientsvanishes (see section 8.18), i.e. unless∣ dx/ds dy/dsA B ∣ =0.At each point in the xy-plane this equation determines a set of curves calledcharacteristic curves (or just characteristics), which thus satisfyB dxds − Ady ds =0,or, multiplying through by ds/dx <strong>and</strong> dividing through by A,dy B(x, y)=dx A(x, y) . (20.41)However, we have already met (20.41) in subsection 20.3.1 on first-order PDEs,where solutions of the <strong>for</strong>m u(x, y) =f(p), where p is some combination of x <strong>and</strong> y,699


PDES: GENERAL AND PARTICULAR SOLUTIONSwere discussed. Comparing (20.41) with (20.12) we see that the characteristics aremerely those curves along which p is constant.Since the partial derivatives ∂u/∂x <strong>and</strong> ∂u/∂y may be evaluated provided theboundary curve C does not lie along a characteristic, defining u(x, y) =φ(s)along C is sufficient to specify the solution to the original problem (equationplus boundary conditions) near the curve C, in terms of a Taylor expansionabout C. There<strong>for</strong>e the characteristics can be considered as the curves alongwhich in<strong>for</strong>mation about the solution u(x, y) ‘propagates’. This is best understoodby using an example.◮Find the general solution ofx ∂u ∂u− 2y∂x ∂y = 0 (20.42)that takes the value 2y +1 on the line x =1betweeny =0<strong>and</strong>y =1.We solved this problem in subsection 20.3.1 <strong>for</strong> the case where u(x, y) takes the value2y +1 along the entire line x = 1. We found then that the general solution to the equation(ignoring boundary conditions) is of the <strong>for</strong>mu(x, y) =f(p) =f(x 2 y),<strong>for</strong> some arbitrary function f. Hence the characteristics of (20.42) are given by x 2 y = cwhere c is a constant; some of these curves are plotted in figure 20.2 <strong>for</strong> various values ofc. Furthermore, we found that the particular solution <strong>for</strong> which u(1,y)=2y +1<strong>for</strong> all ywas given byu(x, y) =2x 2 y +1.In the present case the value of x 2 y is fixed by the boundary conditions only betweeny =0<strong>and</strong>y = 1. However, since the characteristics are curves along which x 2 y, <strong>and</strong> hencef(x 2 y), remains constant, the solution is determined everywhere along any characteristicthat intersects the line segment denoting the boundary conditions. Thus u(x, y) =2x 2 y +1is the particular solution that holds in the shaded region in figure 20.2 (corresponding to0 ≤ c ≤ 1).Outside this region, however, the solution is not precisely specified, <strong>and</strong> any function ofthe <strong>for</strong>mu(x, y) =2x 2 y +1+g(x 2 y)will satisfy both the equation <strong>and</strong> the boundary condition, provided g(p) = 0 <strong>for</strong>0 ≤ p ≤ 1. ◭In the above example the boundary curve was not itself a characteristic <strong>and</strong>furthermore it crossed each characteristic once only. For a general boundary curveC this may not be the case. Firstly, if C is itself a characteristic (or is just a singlepoint) then in<strong>for</strong>mation about the solution cannot ‘propagate’ away from C, <strong>and</strong>so the solution remains unspecified everywhere except on C.The second possibility is that C (although not a characteristic itself) crossessome characteristics more than once, as in figure 20.3. In this case specifying thevalue of u(x, y) along the curve PQdetermines the solution along all the characteristicsthat intersect it. There<strong>for</strong>e, also specifying u(x, y) along QR can overdeterminethe problem solution <strong>and</strong> generally results in there being no solution.700


20.6 CHARACTERISTICS AND THE EXISTENCE OF SOLUTIONSy2c =11−11xy = c/x 2x =1Figure 20.2 The characteristics of equation (20.42). The shaded region showswhere the solution to the equation is defined, given the imposed boundarycondition at x = 1 between y =0<strong>and</strong>y = 1, shown as a bold vertical line.yRPCQxFigure 20.3A boundary curve C that crosses characteristics more than once.20.6.2 Second-order equationsThe concept of characteristics can be extended naturally to second- (<strong>and</strong> higher-)order equations. In this case let us write the general second-order linear PDE(20.19) as(A(x, y) ∂2 u∂x 2 + B(x, y) ∂2 u∂x∂y + C(x, y) ∂2 u∂y 2 = F x, y, u, ∂u∂x , ∂u ). (20.43)∂y701


PDES: GENERAL AND PARTICULAR SOLUTIONSyCdrdydxˆn dsxFigure 20.4point.A boundary curve C <strong>and</strong> its tangent <strong>and</strong> unit normal at a givenFor second-order equations we might expect that relevant boundary conditionswould involve specifying u, or some of its first derivatives, or both, along asuitable set of boundaries bordering or enclosing the region over which a solutionis sought. Three common types of boundary condition occur <strong>and</strong> are associatedwith the names of Dirichlet, Neumann <strong>and</strong> Cauchy. They are as follows.(i) Dirichlet: The value of u is specified at each point of the boundary.(ii) Neumann: The value of ∂u/∂n, thenormal derivative of u, isspecifiedateach point of the boundary. Note that ∂u/∂n = ∇u · ˆn, whereˆn is thenormal to the boundary at each point.(iii) Cauchy: Bothu <strong>and</strong> ∂u/∂n are specified at each point of the boundary.Let us consider <strong>for</strong> the moment the solution of (20.43) subject to the Cauchyboundary conditions, i.e. u <strong>and</strong> ∂u/∂n are specified along some boundary curveC in the xy-plane defined by the parametric equations x = x(s), y = y(s), s beingthe arc length along C (see figure 20.4). Let us suppose that along C we haveu(x, y) =φ(s) <strong>and</strong>∂u/∂n = ψ(s). At any point on C the vector dr = dx i + dy j isa tangent to the curve <strong>and</strong> ˆn ds = dy i − dx j is a vector normal to the curve. Thuson C we have∂u dr≡∇u ·∂s ds = ∂u dx∂x ds + ∂u dy∂y ds = dφ(s) ,ds∂u∂n≡∇u · ˆn =∂u∂xdyds − ∂u∂ydxds = ψ(s).These two equations may then be solved straight<strong>for</strong>wardly <strong>for</strong> the first partialderivatives ∂u/∂x <strong>and</strong> ∂u/∂y along C. Using the chain rule to writedds = dx ∂ds ∂x + dy ∂ds ∂y ,702


20.6 CHARACTERISTICS AND THE EXISTENCE OF SOLUTIONSwe may differentiate the two first derivatives ∂u/∂x <strong>and</strong> ∂u/∂y along the boundaryto obtain the pair of equationsddsdds( ) ∂u∂x( ) ∂u∂y= dx ∂ 2 uds= dxds∂x 2 + dy ∂ 2 uds ∂x∂y ,∂ 2 u∂x∂y + dy ∂ 2 uds ∂y 2 .We may now solve these two equations, together with the original PDE (20.43),<strong>for</strong> the second partial derivatives of u, except where the determinant of theircoefficients equals zero,A B Cdx dy0∣ ds ds∣ =0.∣Exp<strong>and</strong>ing out the determinant,A( ) 2 dy− Bds0( dxdsdxdsMultiplying through by (ds/dx) 2 we obtainAdyds∣)( ) dy+ Cds( ) 2 dx=0.ds( ) dy 2− B dy + C =0, (20.44)dx dxwhich is the ODE <strong>for</strong> the curves in the xy-plane along which the second partialderivatives of u cannot be found.As <strong>for</strong> the first-order case, the curves satisfying (20.44) are called characteristicsof the original PDE. These characteristics have tangents at each point given by(when A ≠0)dydx = B ± √ B 2 − 4AC. (20.45)2AClearly, when the original PDE is hyperbolic (B 2 > 4AC), equation (20.45)defines two families of real curves in the xy-plane; when the equation is parabolic(B 2 =4AC) it defines one family of real curves; <strong>and</strong> when the equation is elliptic(B 2 < 4AC) it defines two families of complex curves. Furthermore, when A,B <strong>and</strong> C are constants, rather than functions of x <strong>and</strong> y, the equations of thecharacteristics will be of the <strong>for</strong>m x + λy = constant, which is reminiscent of the<strong>for</strong>m of solution discussed in subsection 20.3.3.703


PDES: GENERAL AND PARTICULAR SOLUTIONSctx − ct =constant0Lxx + ct =constantFigure 20.5 The characteristics <strong>for</strong> the one-dimensional wave equation. Theshaded region indicates the region over which the solution is determined byspecifying Cauchy boundary conditions at t = 0 on the line segment x =0tox = L.◮Find the characteristics of the one-dimensional wave equation∂ 2 u∂x − 1 ∂ 2 u2 c 2 ∂t =0. 2This is a hyperbolic equation with A =1,B =0<strong>and</strong>C = −1/c 2 . There<strong>for</strong>e from (20.44)the characteristics are given by( ) 2 dx= c 2 ,dt<strong>and</strong> so the characteristics are the straight lines x − ct =constant<strong>and</strong>x + ct =constant.◭The characteristics of second-order PDEs can be considered as the curves alongwhich partial in<strong>for</strong>mation about the solution u(x, y) ‘propagates’. Consider a pointin the space that has the independent variables as its coordinates; unless bothof the two characteristics that pass through the point intersect the curve alongwhich the boundary conditions are specified, the solution will not be determinedat that point. In particular, if the equation is hyperbolic, so that we obtain twofamilies of real characteristics in the xy-plane, then Cauchy boundary conditionspropagate partial in<strong>for</strong>mation concerning the solution along the characteristics,belonging to each family, that intersect the boundary curve C. Thesolutionuis then specified in the region common to these two families of characteristics.For instance, the characteristics of the hyperbolic one-dimensional wave equationin the last example are shown in figure 20.5. By specifying Cauchy boundary704


20.7 UNIQUENESS OF SOLUTIONSEquation type Boundary Conditionshyperbolic open Cauchyparabolic open Dirichlet or Neumannelliptic closed Dirichlet or NeumannTable 20.1 The appropriate boundary conditions <strong>for</strong> different types of partialdifferential equation.conditions u <strong>and</strong> ∂u/∂t on the line segment t =0,x =0toL, the solution isspecified in the shaded region.As in the case of first-order PDEs, however, problems can arise. For example,if <strong>for</strong> a hyperbolic equation the boundary curve intersects any characteristicmore than once then Cauchy conditions along C can overdetermine the problem,resulting in there being no solution. In this case either the boundary curve Cmust be altered, or the boundary conditions on the offending parts of C must berelaxed to Dirichlet or Neumann conditions.The general considerations involved in deciding which boundary conditions areappropriate <strong>for</strong> a particular problem are complex, <strong>and</strong> we do not discuss themany further here. § We merely note that whether the various types of boundarycondition are appropriate (in that they give a solution that is unique, sometimesto within a constant, <strong>and</strong> is well defined) depends upon the type of second-orderequation under consideration <strong>and</strong> on whether the region of solution is boundedby a closed or an open curve (or a surface if there are more than two independentvariables). Note that part of a closed boundary may be at infinity if conditionsare imposed on u or ∂u/∂n there.It may be shown that the appropriate boundary-condition <strong>and</strong> equation-typepairings are as given in table 20.1.For example, Laplace’s equation ∇ 2 u = 0 is elliptic <strong>and</strong> thus requires eitherDirichlet or Neumann boundary conditions on a closed boundary which, as wehave already noted, may be at infinity if the behaviour of u is specified there(most often u or ∂u/∂n → 0 at infinity).20.7 Uniqueness of solutionsAlthough we have merely stated the appropriate boundary types <strong>and</strong> conditions<strong>for</strong> which, in the general case, a PDE has a unique, well-defined solution, sometimesto within an additive constant, it is often important to be able to provethat a unique solution is obtained.§ For a discussion the reader is referred, <strong>for</strong> example, to P. M. Morse <strong>and</strong> H. Feshbach, <strong>Methods</strong> ofTheoretical <strong>Physics</strong>, Part I (New York: McGraw-Hill, 1953), chap. 6.705


PDES: GENERAL AND PARTICULAR SOLUTIONSAs an important example let us consider Poisson’s equation in three dimensions,∇ 2 u(r) =ρ(r), (20.46)with either Dirichlet or Neumann conditions on a closed boundary appropriateto such an elliptic equation; <strong>for</strong> brevity, in (20.46), we have absorbed any physicalconstants into ρ. We aim to show that, to within an unimportant constant, thesolution of (20.46) is unique if either the potential u or its normal derivative∂u/∂n is specified on all surfaces bounding a given region of space (including, ifnecessary, a hypothetical spherical surface of indefinitely large radius on which uor ∂u/∂n is prescribed to have an arbitrarily small value). Stated more <strong>for</strong>mallythis is as follows.Uniqueness theorem. If u is real <strong>and</strong> its first <strong>and</strong> second partial derivatives arecontinuous in a region V <strong>and</strong> on its boundary S, <strong>and</strong> ∇ 2 u = ρ in V <strong>and</strong> eitheru = f or ∂u/∂n = g on S, whereρ, f <strong>and</strong> g are prescribed functions, then u isunique (at least to within an additive constant).◮Prove the uniqueness theorem <strong>for</strong> Poisson’s equation.Let us suppose on the contrary that two solutions u 1 (r)<strong>and</strong>u 2 (r) both satisfy the conditionsgiven above, <strong>and</strong> denote their difference by the function w = u 1 − u 2 . We then have∇ 2 w = ∇ 2 u 1 −∇ 2 u 2 = ρ − ρ =0,so that w satisfies Laplace’s equation in V . Furthermore, since either u 1 = f = u 2 or∂u 1 /∂n = g = ∂u 2 /∂n on S, we must have either w =0or∂w/∂n =0onS.If we now use Green’s first theorem, (11.19), <strong>for</strong> the case where both scalar functionsare taken as w we have ∫[w∇ 2 w +(∇w) · (∇w) ] ∫dV = w ∂wVS ∂n dS.However, either condition, w =0or∂w/∂n = 0, makes the RHS vanish whilst the firstterm on the LHS vanishes since ∇ 2 w =0inV . Thus we are left with∫|∇w| 2 dV =0.VSince |∇w| 2 can never be negative, this can only be satisfied if∇w = 0,i.e. if w, <strong>and</strong> hence u 1 − u 2 , is a constant in V .If Dirichlet conditions are given then u 1 ≡ u 2 on (some part of) S <strong>and</strong> hence u 1 = u 2everywhere in V . For Neumann conditions, however, u 1 <strong>and</strong> u 2 can differ throughout Vby an arbitrary (but unimportant) constant. ◭The importance of this uniqueness theorem lies in the fact that if a solution toPoisson’s (or Laplace’s) equation that fits the given set of Dirichlet or Neumannconditions can be found by any means whatever, then that solution is the correctone, since only one exists. This result is the mathematical justification <strong>for</strong> themethod of images, which is discussed more fully in the next chapter.706


20.8 EXERCISESWe also note that often the same general method, used in the above example<strong>for</strong> proving the uniqueness theorem <strong>for</strong> Poisson’s equation, can be employed toprove the uniqueness (or otherwise) of solutions to other equations <strong>and</strong> boundaryconditions.20.8 Exercises20.1 Determine whether the following can be written as functions of p = x 2 +2y only,<strong>and</strong> hence whether they are solutions of (20.8):(a) x 2 (x 2 − 4) + 4y(x 2 − 2) + 4(y 2 − 1);(b) x 4 +2x 2 y + y 2 ;(c) [x 4 +4x 2 y +4y 2 +4]/[2x 4 + x 2 (8y +1)+8y 2 +2y].20.2 Find partial differential equations satisfied by the following functions u(x, y) <strong>for</strong>all arbitrary functions f <strong>and</strong> all arbitrary constants a <strong>and</strong> b:(a) u(x, y) =f(x 2 − y 2 );(b) u(x, y) =(x − a) 2 +(y − b) 2 ;(c) u(x, y) =y n f(y/x);(d) u(x, y) =f(x + ay).20.3 Solve the following partial differential equations <strong>for</strong> u(x, y) with the boundaryconditions given:(a) x ∂u + xy = u, u =2y on the line x =1;∂x(b) 1 + x ∂u = xu, u(x, 0) = x.∂y20.4 Find the most general solutions u(x, y) of the following equations, consistent withthe boundary conditions stated:(a) y ∂u∂x − x ∂u∂y =0,u(x, 0)=1+sinx;(b) i ∂u∂x =3∂u ∂y , u =(4+3i)x2 on the line x = y;(c)sin x sin y ∂u∂u+cosx cos y∂x ∂y =0,u =cos2y on x + y = π/2;∂u ∂u(d) +2x∂x ∂y =0, u = 2 on the parabola y = x2 .20.5 Find solutions of1 ∂ux ∂x + 1 ∂uy ∂y =0<strong>for</strong> which (a) u(0,y)=y <strong>and</strong> (b) u(1, 1) = 1.20.6 Find the most general solutions u(x, y) of the following equations consistent withthe boundary conditions stated:(a) y ∂u∂x − x ∂u∂y =3x, u = x2 on the line y =0;707


PDES: GENERAL AND PARTICULAR SOLUTIONS(b) y ∂u∂x − x ∂u =3x, u(1, 0) = 2;∂y(c) y 2 ∂u ∂u+ x2∂x ∂y = x2 y 2 (x 3 + y 3 ), no boundary conditions.20.7 Solvesin x ∂u ∂u+cosx∂x ∂y =cosxsubject to (a) u(π/2,y)=0<strong>and</strong>(b)u(π/2,y)=y(y +1).20.8 A function u(x, y) satisfies2 ∂u∂x +3∂u ∂y =10,<strong>and</strong> takes the value 3 on the line y =4x. Evaluate u(2, 4).20.9 If u(x, y) satisfies∂ 2 u∂x − 3 ∂2 u u2 ∂x∂y +2∂2 ∂y =0 2<strong>and</strong> u = −x 2 <strong>and</strong> ∂u/∂y =0<strong>for</strong>y =0<strong>and</strong>allx, find the value of u(0, 1).20.10 Consider the partial differential equation∂ 2 u∂x − 3 ∂2 u u2 ∂x∂y +2∂2 ∂y =0. (∗)2(a) Find the function u(x, y) that satisfies (∗) <strong>and</strong> the boundary condition u =∂u/∂y =1wheny =0<strong>for</strong>allx. Evaluate u(0, 1).(b) In which region of the xy-plane would u be determined if the boundarycondition were u = ∂u/∂y =1wheny =0<strong>for</strong>allx>0?20.11 In those cases in which it is possible to do so, evaluate u(2, 2), where u(x, y) isthe solution of2y ∂u∂x − x ∂u∂y = xy(2y2 − x 2 )that satisfies the (separate) boundary conditions given below.(a) u(x, 1) = x 2 <strong>for</strong> all x.(b) u(x, 1) = x 2 <strong>for</strong> x ≥ 0.(c) u(x, 1) = x 2 <strong>for</strong> 0 ≤ x ≤ 3.(d) u(x, 0) = x <strong>for</strong> x ≥ 0.(e) u(x, 0) = x <strong>for</strong> all x.(f) u(1, √ 10) = 5.(g) u( √ 10, 1) = 5.20.12 Solve6 ∂2 u∂x − 5 ∂2 u2 ∂x∂y + ∂2 u∂y =14, 2subject to u =2x +1<strong>and</strong>∂u/∂y =4− 6x, both on the line y =0.20.13 By changing the independent variables in the previous exercise toξ = x +2y <strong>and</strong> η = x +3y,show that it must be possible to write 14(x 2 +5xy +6y 2 )inthe<strong>for</strong>mf 1 (x +2y)+f 2 (x +3y) − (x 2 + y 2 ),<strong>and</strong> determine the <strong>for</strong>ms of f 1 (z) <strong>and</strong>f 2 (z).708


20.8 EXERCISES20.14 Solve∂ 2 u u∂x∂y +3∂2 = x(2y +3x).∂y2 20.15 Find the most general solution of ∂ 2 u/∂x 2 + ∂ 2 u/∂y 2 = x 2 y 2 .20.16 An infinitely long string on which waves travel at speed c has an initial displacement{sin(πx/a), −a ≤ x ≤ a,y(x) =0, |x| >a.It is released from rest at time t = 0, <strong>and</strong> its subsequent displacement is describedby y(x, t).By expressing the initial displacement as one explicit function incorporatingHeaviside step functions, find an expression <strong>for</strong> y(x, t) atageneraltimet>0. Inparticular, determine the displacement as a function of time (a) at x =0,(b)atx = a, <strong>and</strong>(c)atx = a/2.20.17 The non-relativistic Schrödinger equation (20.7) is similar to the diffusion equationin having different orders of derivatives in its various terms; this precludessolutions that are arbitrary functions of particular linear combinations of variables.However, since exponential functions do not change their <strong>for</strong>ms underdifferentiation, solutions in the <strong>for</strong>m of exponential functions of combinations ofthe variables may still be possible.Consider the Schrödinger equation <strong>for</strong> the case of a constant potential, i.e. <strong>for</strong>a free particle, <strong>and</strong> show that it has solutions of the <strong>for</strong>m A exp(lx+my +nz +λt),where the only requirement is that− 2 (l 2 + m 2 + n 2) = iλ.2mIn particular, identify the equation <strong>and</strong> wavefunction obtained by taking λ as−iE/, <strong>and</strong>l, m <strong>and</strong> n as ip x /,ip y / <strong>and</strong> ip z /, respectively, where E is theenergy <strong>and</strong> p the momentum of the particle; these identifications are essentiallythe content of the de Broglie <strong>and</strong> Einstein relationships.20.18 Like the Schrödinger equation of the previous exercise, the equation describingthe transverse vibrations of a rod,a 4 ∂4 u∂x + ∂2 u4 ∂t =0, 2has different orders of derivatives in its various terms. Show, however, that ithas solutions of exponential <strong>for</strong>m, u(x, t) =A exp(λx + iωt), provided that therelation a 4 λ 4 = ω 2 is satisfied.Use a linear combination of such allowed solutions, expressed as the sum ofsinusoids <strong>and</strong> hyperbolic sinusoids of λx, to describe the transverse vibrations ofa rod of length L clamped at both ends. At a clamped point both u <strong>and</strong> ∂u/∂xmust vanish; show that this implies that cos(λL)cosh(λL) = 1, thus determiningthe frequencies ω at which the rod can vibrate.20.19 An incompressible fluid of density ρ <strong>and</strong> negligible viscosity flows with velocity valong a thin, straight, perfectly light <strong>and</strong> flexible tube, of cross-section A which isheld under tension T . Assume that small transverse displacements u of the tubeare governed by∂ 2 u∂t +2v ∂2 u(v 2 ∂x∂t + 2 − T ) ∂ 2 uρA ∂x =0. 2(a) Show that the general solution consists of a superposition of two wave<strong>for</strong>mstravelling with different speeds.709


PDES: GENERAL AND PARTICULAR SOLUTIONS(b) The tube initially has a small transverse displacement u = a cos kx <strong>and</strong> issuddenly released from rest. Find its subsequent motion.20.20 A sheet of material of thickness w, specific heat capacity c <strong>and</strong> thermal conductivityk is isolated in a vacuum, but its two sides are exposed to fluxes ofradiant heat of strengths J 1 <strong>and</strong> J 2 . Ignoring short-term transients, show that thetemperature difference between its two surfaces is steady at (J 2 − J 1 )w/2k, whilsttheir average temperature increases at a rate (J 2 + J 1 )/cw.20.21 In an electrical cable of resistance R <strong>and</strong> capacitance C, each per unit length,voltage signals obey the equation ∂ 2 V/∂x 2 = RC∂V/∂t. This has solutions of the<strong>for</strong>m given in (20.36) <strong>and</strong> also of the <strong>for</strong>m V = Ax + D.(a) Find a combination of these that represents the situation after a steadyvoltage V 0 is applied at x =0attimet =0.(b) Obtain a solution describing the propagation of the voltage signal resultingfrom the application of the signal V = V 0 <strong>for</strong> 0


20.9 HINTS AND ANSWERS20.25 The Klein–Gordon equation (which is satisfied by the quantum-mechanical wavefunctionΦ(r) of a relativistic spinless particle of non-zero mass m) is∇ 2 Φ − m 2 Φ=0.Show that the solution <strong>for</strong> the scalar field Φ(r) inanyvolumeV bounded byasurfaceS is unique if either Dirichlet or Neumann boundary conditions arespecified on S.20.9 Hints <strong>and</strong> answers20.1 (a) Yes, p 2 − 4p − 4; (b) no, (p − y) 2 ;(c)yes,(p 2 +4)/(2p 2 + p).20.3 Each equation is effectively an ordinary differential equation, but with a functionof the non-integrated variable as the constant of integration;(a) u = xy(2 − ln x); (b) u = x −1 (1 − e y )+xe y .20.5 (a) (y 2 − x 2 ) 1/2 ;(b)1+f(y 2 − x 2 ), where f(0) = 0.20.7 u = y + f(y − ln(sin x)); (a) u =ln(sinx); (b) u = y +[y − ln(sin x)] 2 .20.9 General solution is u(x, y) =f(x + y) +g(x + y/2). Show that 2p = −g ′ (p)/2,<strong>and</strong> hence g(p) =k − 2p 2 , whilst f(p) =p 2 − k, leading to u(x, y) =−x 2 + y 2 /2;u(0, 1) = 1/2.20.11 p = x 2 +2y 2 ; u(x, y) =f(p)+x 2 y 2 /2.(a) u(x, y) =(x 2 +2y 2 + x 2 y 2 − 2)/2; u(2, 2) = 13. The line y = 1 cuts eachcharacteristic in zero or two distinct points, but this causes no difficulty withthe given boundary conditions.(b) As in (a).(c) The solution is defined over the space between the ellipses p =2<strong>and</strong>p = 11;(2, 2) lies on p = 12, <strong>and</strong> so u(2, 2) is undetermined.(d) u(x, y) =(x 2 +2y 2 ) 1/2 + x 2 y 2 /2; u(2, 2) = 8 + √ 12.(e) The line y = 0 cuts each characteristic in two distinct points. No differentiable<strong>for</strong>m of f(p) givesf(±a) =±a respectively, <strong>and</strong> so there is no solution.(f) The solution is only specified on p = 21, <strong>and</strong> so u(2, 2) is undetermined.(g) The solution is specified on p = 12, <strong>and</strong> so u(2, 2) = 5 + 1 (4)(4) = 13.220.13 The equation becomes ∂ 2 f/∂ξ∂η = −14, with solution f(ξ,η) =f(ξ)+g(η)−14ξη,which can be compared with the answer from the previous question; f 1 (z) =10z 2<strong>and</strong> f 2 (z) =5z 2 .20.15 u(x, y) =f(x + iy)+g(x − iy)+(1/12)x 4 (y 2 − (1/15)x 2 ). In the last term, x <strong>and</strong>y may be interchanged. There are (infinitely) many other possibilities <strong>for</strong> thespecific PI, e.g. [ 15x 2 y 2 (x 2 + y 2 ) − (x 6 + y 6 )]/360.20.17 E = p 2 /(2m), the relationship between energy <strong>and</strong> momentum <strong>for</strong> a nonrelativisticparticle; u(r,t)=A exp[i(p · r − Et)/], a plane wave of wave numberk = p/ <strong>and</strong> angular frequency ω = E/ travelling in the direction p/p.20.19 (a) c = v ± α where α 2 = T/ρA;(b) u(x,[t) =a cos[k(x − vt)] cos(kαt) − (va/α)sin[k(x − vt)] sin(kαt).20.21 (a) V 0 1 − (2/ √ π) ∫ ]12 x(CR/t)1/2 exp(−ν 2 ) dν ;(b) consider the input as equivalent to V 0 applied at t = 0 <strong>and</strong> continued <strong>and</strong>−V 0 applied at t = T <strong>and</strong> continued;V (x, t) = 2V ∫ 10 2 x[CR/(t−T )] 1/2√ exp ( −ν 2) dν;π 12 x(CR/t) 1/2(c) For t ≫ T , maximum at x =[2t/(CR)] 1/2 with value V 0T exp(− 1 ) 2.(2π) 1/2 t711


PDES: GENERAL AND PARTICULAR SOLUTIONS20.23 (a) Parabolic, open, Dirichlet u(x, 0) given, Neumann ∂u/∂x =0atx = ±L/2<strong>for</strong> all t;(b) elliptic, closed, Dirichlet;(c) elliptic, closed, Neumann ∂u/∂n = σ/ɛ 0 ;(d) hyperbolic, open, Cauchy.20.25 Follow an argument similar to that in section 20.7 <strong>and</strong> argue that the additionalterm ∫ m 2 |w| 2 dV must be zero, <strong>and</strong> hence that w = 0 everywhere.712


21Partial differential equations:separation of variables <strong>and</strong> othermethodsIn the previous chapter we demonstrated the methods by which general solutionsof some partial differential equations (PDEs) may be obtained in terms ofarbitrary functions. In particular, solutions containing the independent variablesin definite combinations were sought, thus reducing the effective number of them.In the present chapter we begin by taking the opposite approach, namely thatof trying to keep the independent variables as separate as possible, using themethod of separation of variables. We then consider integral trans<strong>for</strong>m methodsby which one of the independent variables may be eliminated, at least fromdifferential coefficients. Finally, we discuss the use of Green’s functions in solvinginhomogeneous problems.21.1 Separation of variables: the general methodSuppose we seek a solution u(x, y, z, t) to some PDE (expressed in Cartesiancoordinates). Let us attempt to obtain one that has the product <strong>for</strong>m §u(x, y, z, t) =X(x)Y (y)Z(z)T (t). (21.1)A solution that has this <strong>for</strong>m is said to be separable in x, y, z <strong>and</strong> t, <strong>and</strong> seekingsolutions of this <strong>for</strong>m is called the method of separation of variables.As simple examples we may observe that, of the functions(i) xyz 2 sin bt, (ii) xy + zt, (iii) (x 2 + y 2 )z cos ωt,(i) is completely separable, (ii) is inseparable in that no single variable can beseparated out from it <strong>and</strong> written as a multiplicative factor, whilst (iii) is separablein z <strong>and</strong> t but not in x <strong>and</strong> y.§ It should be noted that the conventional use here of upper-case (capital) letters to denote thefunctions of the corresponding lower-case variable is intended to enable an easy correspondencebetween a function <strong>and</strong> its argument to be made.713


PDES: SEPARATION OF VARIABLES AND OTHER METHODSWhen seeking PDE solutions of the <strong>for</strong>m (21.1), we are requiring not thatthere is no connection at all between the functions X, Y , Z <strong>and</strong> T (<strong>for</strong> example,certain parameters may appear in two or more of them), but only that X doesnot depend upon y, z, t, thatY does not depend on x, z, t, <strong>and</strong>soon.For a general PDE it is likely that a separable solution is impossible, butcertainly some common <strong>and</strong> important equations do have useful solutions ofthis <strong>for</strong>m, <strong>and</strong> we will illustrate the method of solution by studying the threedimensionalwave equation∇ 2 u(r) = 1 ∂ 2 u(r)c 2 ∂t 2 . (21.2)We will work in Cartesian coordinates <strong>for</strong> the present <strong>and</strong> assume a solutionof the <strong>for</strong>m (21.1); the solutions in alternative coordinate systems, e.g. sphericalor cylindrical polars, are considered in section 21.3. Expressed in Cartesiancoordinates (21.2) takes the <strong>for</strong>msubstituting (21.1) gives∂ 2 u∂x 2 + ∂2 u∂y 2 + ∂2 u∂z 2 = 1 c 2 ∂ 2 u∂t 2 ; (21.3)d 2 Xdx 2 YZT + X d2 Ydy 2 ZT + XY d2 Zdz 2 T = 1 c 2 XY Z d2 Tdt 2 ,which can also be written asX ′′ YZT + XY ′′ ZT + XY Z ′′ T = 1 c 2 XY ZT′′ , (21.4)whereineachcasetheprimesrefertotheordinary derivative with respect to theindependent variable upon which the function depends. This emphasises the factthat each of the functions X, Y , Z <strong>and</strong> T has only one independent variable <strong>and</strong>thus its only derivative is its total derivative. For the same reason, in each termin (21.4) three of the four functions are unaltered by the partial differentiation<strong>and</strong> behave exactly as constant multipliers.If we now divide (21.4) throughout by u = XY ZT we obtainX ′′X + Y ′′Y + Z ′′Z = 1 T ′′c 2 T . (21.5)This <strong>for</strong>m shows the particular characteristic that is the basis of the method ofseparation of variables, namely that of the four terms the first is a function of xonly, the second of y only, the third of z only <strong>and</strong> the RHS a function of t only<strong>and</strong> yet there is an equation connecting them. This can only be so <strong>for</strong> all x, y, z<strong>and</strong> t if each of the terms does not in fact, despite appearances, depend upon thecorresponding independent variable but is equal to a constant, the four constantsbeing such that (21.5) is satisfied.714


21.1 SEPARATION OF VARIABLES: THE GENERAL METHODSince there is only one equation to be satisfied <strong>and</strong> four constants involved,there is considerable freedom in the values they may take. For the purposes ofour illustrative example let us make the choice of −l 2 , −m 2 , −n 2 ,<strong>for</strong>thefirstthree constants. The constant associated with c −2 T ′′ /T must then have the value−µ 2 = −(l 2 + m 2 + n 2 ).Having recognised that each term of (21.5) is individually equal to a constant(or parameter), we can now replace (21.5) by four separate ordinary differentialequations (ODEs):X ′′X = Y ′′−l2 ,Y = Z ′′−m2 ,Z = 1 T ′′−n2 ,c 2 T = −µ2 . (21.6)The important point to notice is not the simplicity of the equations (21.6) (thecorresponding ones <strong>for</strong> a general PDE are usually far from simple) but that, bythe device of assuming a separable solution, a partial differential equation (21.3),containing derivatives with respect to the four independent variables all in oneequation, has been reduced to four separate ordinary differential equations (21.6).The ordinary equations are connected through four constant parameters thatsatisfy an algebraic relation. These constants are called separation constants.The general solutions of the equations (21.6) can be deduced straight<strong>for</strong>wardly<strong>and</strong> areX(x) =A exp(ilx)+B exp(−ilx),Y (y) =C exp(imy)+D exp(−imy),Z(z) =E exp(inz)+F exp(−inz),T (t) =G exp(icµt)+H exp(−icµt),(21.7)where A, B,...,H are constants, which may be determined if boundary condtionsare imposed on the solution. Depending on the geometry of the problem <strong>and</strong>any boundary conditions, it is sometimes more appropriate to write the solutions(21.7) in the alternative <strong>for</strong>mX(x) =A ′ cos lx + B ′ sin lx,Y (y) =C ′ cos my + D ′ sin my,Z(z) =E ′ cos nz + F ′ sin nz,T (t) =G ′ cos(cµt)+H ′ sin(cµt),(21.8)<strong>for</strong> some different set of constants A ′ ,B ′ ,...,H ′ . Clearly the choice of how bestto represent the solution depends on the problem being considered.As an example, suppose that we take as particular solutions the four functionsX(x) =exp(ilx),Z(z) =exp(inz),Y(y) =exp(imy),T(t) =exp(−icµt).715


PDES: SEPARATION OF VARIABLES AND OTHER METHODSThis gives a particular solution of the original PDE (21.3)u(x, y, z, t) =exp(ilx) exp(imy) exp(inz) exp(−icµt)= exp[i(lx + my + nz − cµt)],which is a special case of the solution (20.33) obtained in the previous chapter<strong>and</strong> represents a plane wave of unit amplitude propagating in a direction givenby the vector with components l,m,n in a Cartesian coordinate system. In theconventional notation of wave theory, l, m <strong>and</strong> n are the components of thewave-number vector k, whose magnitude is given by k =2π/λ, whereλ is thewavelength of the wave; cµ is the angular frequency ω of the wave. This givesthe equation in the <strong>for</strong>mu(x, y, z, t) =exp[i(k x x + k y y + k z z − ωt)]= exp[i(k · r − ωt)],<strong>and</strong> makes the exponent dimensionless.The method of separation of variables can be applied to many commonlyoccurring PDEs encountered in physical applications.◮Use the method of separation of variables to obtain <strong>for</strong> the one-dimensional diffusionequationκ ∂2 u∂x = ∂u2 ∂t , (21.9)a solution that tends to zero as t →∞<strong>for</strong> all x.Here we have only two independent variables x <strong>and</strong> t <strong>and</strong> we there<strong>for</strong>e assume a solutionof the <strong>for</strong>mu(x, t) =X(x)T (t).Substituting this expression into (21.9) <strong>and</strong> dividing through by u = XT (<strong>and</strong> also by κ)we obtainX ′′X = T ′κT .Now, arguing exactly as above that the LHS is a function of x only<strong>and</strong>theRHSisafunction of t only, we conclude that each side must equal a constant, which, anticipatingthe result <strong>and</strong> noting the imposed boundary condition, we will take as −λ 2 . This gives ustwo ordinary equations,X ′′ + λ 2 X =0, (21.10)T ′ + λ 2 κT =0, (21.11)which have the solutionsX(x) =A cos λx + B sin λx,T (t) =C exp(−λ 2 κt).Combining these to give the assumed solution u = XT yields (absorbing the constant Cinto A <strong>and</strong> B)u(x, t) =(A cos λx + B sin λx)exp(−λ 2 κt). (21.12)716


21.2 SUPERPOSITION OF SEPARATED SOLUTIONSIn order to satisfy the boundary condition u → 0ast →∞, λ 2 κ must be > 0. Since κis real <strong>and</strong> > 0, this implies that λ is a real non-zero number <strong>and</strong> that the solution issinusoidal in x <strong>and</strong> is not a disguised hyperbolic function; this was our reason <strong>for</strong> choosingthe separation constant as −λ 2 . ◭As a final example we consider Laplace’s equation in Cartesian coordinates;this may be treated in a similar manner.◮Use the method of separation of variables to obtain a solution <strong>for</strong> the two-dimensionalLaplace equation,∂ 2 u∂x + ∂2 u=0. (21.13)2 ∂y2 If we assume a solution of the <strong>for</strong>m u(x, y) =X(x)Y (y) then, following the above method,<strong>and</strong> taking the separation constant as λ 2 , we findX ′′ = λ 2 X, Y ′′ = −λ 2 Y.Taking λ 2 as > 0, the general solution becomesu(x, y) =(A cosh λx + B sinh λx)(C cos λy + D sin λy). (21.14)An alternative <strong>for</strong>m, in which the exponentials are written explicitly, may be useful <strong>for</strong>other geometries or boundary conditions:u(x, y) =[A exp λx + B exp(−λx)](C cos λy + D sin λy), (21.15)with different constants A <strong>and</strong> B.If λ 2 < 0 then the roles of x <strong>and</strong> y interchange. The particular combination of sinusoidal<strong>and</strong> hyperbolic functions <strong>and</strong> the values of λ allowed will be determined by the geometricalproperties of any specific problem, together with any prescribed or necessary boundaryconditions. ◭We note here that a particular case of the solution (21.14) links up with the‘combination’ result u(x, y) =f(x + iy) of the previous chapter (equations (20.24)<strong>and</strong> following), namely that if A = B <strong>and</strong> D = iC then the solution is the sameas f(p) =AC exp λp with p = x + iy.21.2 Superposition of separated solutionsIt will be noticed in the previous two examples that there is considerable freedomin the values of the separation constant λ, the only essential requirement beingthat λ has the same value in both parts of the solution, i.e. the part dependingon x <strong>and</strong> the part depending on y (or t). This is a general feature <strong>for</strong> solutionsin separated <strong>for</strong>m, which, if the original PDE has n independent variables, willcontain n − 1 separation constants. All that is required in general is that weassociate the correct function of one independent variable with the appropriatefunctions of the others, the correct function being the one with the same valuesof the separation constants.If the original PDE is linear (as are the Laplace, Schrödinger, diffusion <strong>and</strong>wave equations) then mathematically acceptable solutions can be <strong>for</strong>med by717


PDES: SEPARATION OF VARIABLES AND OTHER METHODSsuperposing solutions corresponding to different allowed values of the separationconstants. To take a two-variable example: ifu λ1 (x, y) =X λ1 (x)Y λ1 (y)is a solution of a linear PDE obtained by giving the separation constant the valueλ 1 , then the superpositionu(x, y) =a 1 X λ1 (x)Y λ1 (y)+a 2 X λ2 (x)Y λ2 (y)+···= ∑ a i X λi (x)Y λi (y)i(21.16)is also a solution <strong>for</strong> any constants a i , provided that the λ i are the allowed valuesof the separation constant λ given the imposed boundary conditions. Note thatif the boundary conditions allow any of the separation constants to be zero thenthe <strong>for</strong>m of the general solution is normally different <strong>and</strong> must be deduced byreturning to the separated ordinary differential equations. We will encounter thisbehaviour in section 21.3.The value of the superposition approach is that a boundary condition, say thatu(x, y) takes a particular <strong>for</strong>m f(x) wheny = 0, might be met by choosing theconstants a i such thatf(x) = ∑ a i X λi (x)Y λi (0).iIn general, this will be possible provided that the functions X λi (x) <strong>for</strong>m a completeset – as do the sinusoidal functions of Fourier series or the spherical harmonicsdiscussed in subsection 18.3.◮A semi-infinite rectangular metal plate occupies the region 0 ≤ x ≤∞<strong>and</strong> 0 ≤ y ≤ b inthe xy-plane. The temperature at the far end of the plate <strong>and</strong> along its two long sides isfixed at 0 ◦ C. If the temperature of the plate at x =0is also fixed <strong>and</strong> is given by f(y), findthe steady-state temperature distribution u(x,y) of the plate. Hence find the temperaturedistribution if f(y) =u 0 ,whereu 0 is a constant.The physical situation is illustrated in figure 21.1. With the notation we have used severaltimes be<strong>for</strong>e, the two-dimensional heat diffusion equation satisfied by the temperatureu(x, y, t) is( )∂ 2 uκ∂x + ∂2 u= ∂u2 ∂y 2 ∂t ,with κ = k/(sρ). In this case, however, we are asked to find the steady-state temperature,which corresponds to ∂u/∂t = 0, <strong>and</strong> so we are led to consider the (two-dimensional)Laplace equation∂ 2 u∂x + ∂2 u2 ∂y =0. 2We saw that assuming a separable solution of the <strong>for</strong>m u(x, y) =X(x)Y (y) led tosolutions such as (21.14) or (21.15), or equivalent <strong>for</strong>ms with x <strong>and</strong> y interchanged. Inthe current problem we have to satisfy the boundary conditions u(x, 0)=0=u(x, b) <strong>and</strong>so a solution that is sinusoidal in y seems appropriate. Furthermore, since we requireu(∞,y) = 0 it is best to write the x-dependence of the solution explicitly in terms of718


21.2 SUPERPOSITION OF SEPARATED SOLUTIONSybu =0u = f(y)u → 00 u =0xA semi-infinite metal plate whose edges are kept at fixed tem-Figure 21.1peratures.exponentials rather than of hyperbolic functions. We there<strong>for</strong>e write the separable solutionin the <strong>for</strong>m (21.15) asu(x, y) =[A exp λx + B exp(−λx)](C cos λy + D sin λy).Applying the boundary conditions, we see firstly that u(∞,y) = 0 implies A =0ifwetake λ>0. Secondly, since u(x, 0) = 0 we may set C = 0, which, if we absorb the constantD into B, leaves us withu(x, y) =B exp(−λx)sinλy.But, using the condition u(x, b) = 0, we require sin λb =0<strong>and</strong>soλ must be equal to nπ/b,where n is any positive integer.Using the principle of superposition (21.16), the general solution satisfying the givenboundary conditions can there<strong>for</strong>e be written∞∑u(x, y) = B n exp(−nπx/b) sin(nπy/b), (21.17)n=1<strong>for</strong> some constants B n . Notice that in the sum in (21.17) we have omitted negative values ofn since they would lead to exponential terms that diverge as x →∞.Then = 0 term is alsoomitted since it is identically zero. Using the remaining boundary condition u(0,y)=f(y)we see that the constants B n must satisfy∞∑f(y) = B n sin(nπy/b). (21.18)n=1This is clearly a Fourier sine series expansion of f(y) (see chapter 12). For (21.18) tohold, however, the continuation of f(y) outside the region 0 ≤ y ≤ b must be an oddperiodic function with period 2b (see figure 21.2). We also see from figure 21.2 that ifthe original function f(y) does not equal zero at either of y =0<strong>and</strong>y = b then itscontinuation has a discontinuity at the corresponding point(s); nevertheless, as discussedin chapter 12, the Fourier series will converge to the mid-points of these jumps <strong>and</strong> hencetend to zero in this case. If, however, the top <strong>and</strong> bottom edges of the plate were held notat 0 ◦ C but at some other non-zero temperature, then, in general, the final solution wouldpossess discontinuities at the corners x =0,y =0<strong>and</strong>x =0,y = b.Bearing in mind these technicalities, the coefficients B n in (21.18) are given byB n = 2 ∫ b ( nπy)f(y)sin dy. (21.19)bb0719


PDES: SEPARATION OF VARIABLES AND OTHER METHODSf(y)−b 0 b yFigure 21.2The continuation of f(y) <strong>for</strong> a Fourier sine series.There<strong>for</strong>e, if f(y) =u 0 (i.e. the temperature of the side at x = 0 is constant along itslength), (21.19) becomes∫ bB n = 2 b 0[= − 2u 0bThere<strong>for</strong>e the required solution isu 0 sin( nπy)dyb) ] bb( nπynπ cos b= − 2u 0nπ [(−1)n − 1] =u(x, y) = ∑ n odd4u(0nπ exp0{4u0 /nπ <strong>for</strong> n odd,0 <strong>for</strong> n even.− nπxb)sin( nπy). ◭bIn the above example the boundary conditions meant that one term in eachpart of the separable solution could be immediately discarded, making the problemmuch easier to solve. Sometimes, however, a little ingenuity is required inwriting the separable solution in such a way that certain parts can be neglectedimmediately.◮Suppose that the semi-infinite rectangular metal plate in the previous example is replacedby one that in the x-direction has finite length a. The temperature of the right-h<strong>and</strong> edgeis fixed at 0 ◦ C <strong>and</strong> all other boundary conditions remain as be<strong>for</strong>e. Find the steady-statetemperature in the plate.As in the previous example, the boundary conditions u(x, 0) = 0 = u(x, b) suggest a solutionthat is sinusoidal in y. In this case, however, we require u =0onx = a (rather than atinfinity) <strong>and</strong> so a solution in which the x-dependence is written in terms of hyperbolicfunctions, such as (21.14), rather than exponentials is more appropriate. Moreover, sincethe constants in front of the hyperbolic functions are, at this stage, arbitrary, we maywrite the separable solution in the most convenient way that ensures that the conditionu(a, y) = 0 is straight<strong>for</strong>wardly satisfied. We there<strong>for</strong>e writeu(x, y) =[A cosh λ(a − x)+B sinh λ(a − x)](C cos λy + D sin λy).Now the condition u(a, y) = 0 is easily satisfied by setting A = 0. As be<strong>for</strong>e theconditions u(x, 0)=0=u(x, b) implyC =0<strong>and</strong>λ = nπ/b <strong>for</strong> integer n. Superposing the720


21.2 SUPERPOSITION OF SEPARATED SOLUTIONSsolutions <strong>for</strong> different n we then obtainu(x, y) =∞∑B n sinh[nπ(a − x)/b] sin(nπy/b), (21.20)n=1<strong>for</strong> some constants B n . We have omitted negative values of n in the sum (21.20) since therelevant terms are already included in those obtained <strong>for</strong> positive n. Again the n =0termis identically zero. Using the final boundary condition u(0,y)=f(y) as above we find thatthe constants B n must satisfyf(y) =∞∑B n sinh(nπa/b) sin(nπy/b),n=1<strong>and</strong>, remembering the caveats discussed in the previous example, the B n are there<strong>for</strong>e givenby∫ b2B n =f(y)sin(nπy/b) dy. (21.21)b sinh(nπa/b) 0For the case where f(y) =u 0 , following the working of the previous example gives(21.21) asB n =4u 0nπ sinh(nπa/b)<strong>for</strong> n odd, B n =0 <strong>for</strong>n even. (21.22)The required solution is thusu(x, y) = ∑ 4u 0nπ sinh(nπa/b) sinh[nπ(a − x)/b]sin( nπy/b ) .n oddWe note that, as required, in the limit a →∞this solution tends to the solution of theprevious example. ◭Often the principle of superposition can be used to write the solution toproblems with more complicated boundary conditions as the sum of solutions toproblems that each satisfy only some part of the boundary condition but whenadded togther satisfy all the conditions.◮Find the steady-state temperature in the (finite) rectangular plate of the previous example,subject to the boundary conditions u(x, b) =0, u(a, y) =0<strong>and</strong> u(0,y)=f(y) as be<strong>for</strong>e, butnow, in addition, u(x, 0) = g(x).Figure 21.3(c) shows the imposed boundary conditions <strong>for</strong> the metal plate. Although wecould find a solution to this problem using the methods presented above, we can arrive atthe answer almost immediately by using the principle of superposition <strong>and</strong> the result ofthe previous example.Let us suppose the required solution u(x, y) is made up of two parts:u(x, y) =v(x, y)+w(x, y),where v(x, y) is the solution satisfying the boundary conditions shown in figure 21.3(a),721


PDES: SEPARATION OF VARIABLES AND OTHER METHODSyyb0b0f(y) 0000axg(x)ax(a)(b)yb0f(y)0g(x)ax(c)Figure 21.3Superposition of boundary conditions <strong>for</strong> a metal plate.whilst w(x, y) is the solution satisfying the boundary conditions in figure 21.3(b). It is clearthat v(x, y) is simply given by the solution to the previous example,v(x, y) = ∑ [ nπ(a − x)B n sinhbn odd]sin( nπy),bwhere B n is given by (21.21). Moreover, by symmetry, w(x, y) must be of the same <strong>for</strong>m asv(x, y) but with x <strong>and</strong> a interchanged with y <strong>and</strong> b, respectively, <strong>and</strong> with f(y) in (21.21)replaced by g(x). There<strong>for</strong>e the required solution can be written down immediately withoutfurther calculation asu(x, y) = ∑ [ nπ(a − x)n sinhn oddBb]sinthe B n being given by (21.21) <strong>and</strong> the C n byC n =2a sinh(nπb/a)( nπy)+ ∑ n sinhbn oddC∫ a0g(x)sin(nπx/a) dx.[ nπ(b − y)a] ( nπx)sin ,aClearly, this method may be extended to cases in which three or four sides of the platehave non-zero boundary conditions. ◭As a final example of the usefulness of the principle of superposition we nowconsider a problem that illustrates how to deal with inhomogeneous boundaryconditions by a suitable change of variables.722


21.2 SUPERPOSITION OF SEPARATED SOLUTIONS◮A bar of length L is initially at a temperature of 0 ◦ C. One end of the bar (x =0)is heldat 0 ◦ C <strong>and</strong> the other is supplied with heat at a constant rate per unit area of H. Findthetemperature distribution within the bar after a time t.With our usual notation, the heat diffusion equation satisfied by the temperature u(x, t) isκ ∂2 u∂x = ∂u2 ∂t ,with κ = k/(sρ), where k is the thermal conductivity of the bar, s is its specific heatcapacity <strong>and</strong> ρ is its density.The boundary conditions can be written as∂u(L, t)u(x, 0) = 0, u(0,t)=0,= H ∂x k ,the last of which is inhomogeneous. In general, inhomogeneous boundary conditions cancause difficulties <strong>and</strong> it is usual to attempt a trans<strong>for</strong>mation of the problem into anequivalent homogeneous one. To this end, let us assume that the solution to our problemtakes the <strong>for</strong>mu(x, t) =v(x, t)+w(x),where the function w(x) is to be suitably determined. In terms of v <strong>and</strong> w the problembecomes( )∂ 2 vκ∂x + d2 w= ∂v2 dx 2 ∂t ,v(x, 0) + w(x) =0,v(0,t)+w(0) = 0,∂v(L, t)∂x+ dw(L)dx= H k .There are several ways of choosing w(x) so as to make the new problem straight<strong>for</strong>ward.Using some physical insight, however, it is clear that ultimately (at t = ∞), when alltransients have died away, the end x = L will attain a temperature u 0 such that ku 0 /L = H<strong>and</strong> there will be a constant temperature gradient u(x, ∞) =u 0 x/L. We there<strong>for</strong>e choosew(x) = Hxk .Since the second derivative of w(x) is zero, v satisfies the diffusion equation <strong>and</strong> theboundary conditions on v are now given byv(x, 0) = − Hxk , v(0,t)=0, ∂v(L, t)=0,∂xwhich are homogeneous in x.From (21.12) a separated solution <strong>for</strong> the one-dimensional diffusion equation isv(x, t) =(A cos λx + B sin λx)exp(−λ 2 κt),corresponding to a separation constant −λ 2 .Ifwerestrictλ to be real then all thesesolutions are transient ones decaying to zero as t →∞. These are just what is required toadd to w(x) to give the correct solution as t →∞. In order to satisfy v(0,t) = 0, however,we require A = 0. Furthermore, since∂v∂x = B exp(−λ2 κt)λ cos λx,723


PDES: SEPARATION OF VARIABLES AND OTHER METHODSf(x)−L0Lx−HL/kFigure 21.4 The appropriate continuation <strong>for</strong> a Fourier series containingonly sine terms.in order to satisfy ∂v(L, t)/∂x =0werequirecosλL =0,<strong>and</strong>soλ is restricted to thevaluesλ = nπ2L ,where n is an odd non-negative integer, i.e. n =1, 3, 5,... .Thus, to satisfy the boundary condition v(x, 0) = −Hx/k, we must have∑ ( nπx)B n sin = − Hx2L k ,n oddin the range x =0tox = L. In this case we must be more careful about the continuationof the function −Hx/k, <strong>for</strong> which the Fourier sine series is required. We want a series thatis odd in x (sine terms only) <strong>and</strong> continuous as x =0<strong>and</strong>x = L (no discontinuities, sincethe series must converge at the end-points). This leads to a continuation of the functionas shown in figure 21.4, with a period of L ′ =4L. Following the discussion of section 12.3,since this continuation is odd about x = 0 <strong>and</strong> even about x = L ′ /4=L it can indeed beexpressed as a Fourier sine series containing only odd-numbered terms.The corresponding Fourier series coefficients are found to beB n = −8HL (−1) (n−1)/2kπ 2 n 2<strong>and</strong> thus the final <strong>for</strong>mula <strong>for</strong> u(x, t) isu(x, t) = Hxk− 8HLkπ 2∑n odd(−1) (n−1)/2n 2sin<strong>for</strong> n odd,( nπx) ( )exp − kn2 π 2 t,2L 4L 2 sρgiving the temperature <strong>for</strong> all positions 0 ≤ x ≤ L <strong>and</strong> <strong>for</strong> all times t ≥ 0. ◭We note that in all the above examples the boundary conditions restricted theseparation constant(s) to an infinite number of discrete values, usually integers.If, however, the boundary conditions allow the separation constant(s) λ to takea continuum of values then the summation in (21.16) is replaced by an integralover λ. This is discussed further in connection with integral trans<strong>for</strong>m methodsin section 21.4.724


21.3 SEPARATION OF VARIABLES IN POLAR COORDINATES21.3 Separation of variables in polar coordinatesSo far we have considered the solution of PDEs only in Cartesian coordinates,but many systems in two <strong>and</strong> three dimensions are more naturally expressedin some <strong>for</strong>m of polar coordinates, in which full advantage can be taken ofany inherent symmetries. For example, the potential associated with an isolatedpoint charge has a very simple expression, q/(4πɛ 0 r), when polar coordinates areused, but involves all three coordinates <strong>and</strong> square roots when Cartesians areemployed. For these reasons we now turn to the separation of variables in planepolar, cylindrical polar <strong>and</strong> spherical polar coordinates.Most of the PDEs we have considered so far have involved the operator ∇ 2 ,e.g.the wave equation, the diffusion equation, Schrödinger’s equation <strong>and</strong> Poisson’sequation (<strong>and</strong> of course Laplace’s equation). It is there<strong>for</strong>e appropriate that werecall the expressions <strong>for</strong> ∇ 2 when expressed in polar coordinate systems. Fromchapter 10, in plane polars, cylindrical polars <strong>and</strong> spherical polars, respectively,we have∇ 2 = 1 (∂ρ ∂ )+ 1 ∂ 2ρ ∂ρ ∂ρ ρ 2 ∂φ 2 , (21.23)∇ 2 = 1 (∂ρ ∂ )+ 1 ∂ 2ρ ∂ρ ∂ρ ρ 2 ∂φ 2 + ∂2∂z 2 , (21.24)∇ 2 = 1 (∂r 2 r 2 ∂ )+ 1 (∂∂r ∂r r 2 sin θ ∂ )1 ∂ 2+sin θ ∂θ ∂θ r 2 sin 2 θ ∂φ 2 . (21.25)Of course the first of these may be obtained from the second by taking z to beidentically zero.21.3.1 Laplace’s equation in polar coordinatesThe simplest of the equations containing ∇ 2 is Laplace’s equation,∇ 2 u(r) =0. (21.26)Since it contains most of the essential features of the other more complicatedequations, we will consider its solution first.Laplace’s equation in plane polarsSuppose that we need to find a solution of (21.26) that has a prescribed behaviouron the circle ρ = a (e.g. if we are finding the shape taken up by a circular drumskinwhen its rim is slightly de<strong>for</strong>med from being planar). Then we may seek solutionsof (21.26) that are separable in ρ <strong>and</strong> φ (measured from some arbitrary radiusas φ = 0) <strong>and</strong> hope to accommodate the boundary condition by examining thesolution <strong>for</strong> ρ = a.725


PDES: SEPARATION OF VARIABLES AND OTHER METHODSThus, writing u(ρ, φ) =P (ρ)Φ(φ) <strong>and</strong> using the expression (21.23), Laplace’sequation (21.26) becomes(Φ ∂ρ ∂P )+ P ∂ 2 Φρ ∂ρ ∂ρ ρ 2 ∂φ 2 =0.Now, employing the same device as previously, that of dividing through byu = P Φ <strong>and</strong> multiplying through by ρ 2 , results in the separated equation)ρP∂∂ρ(ρ ∂P∂ρ+ 1 ∂ 2 ΦΦ ∂φ 2 =0.Following our earlier argument, since the first term on the RHS is a function ofρ only, whilst the second term depends only on φ, we obtain the two ordinaryequationsρP(dρ dP )= n 2 , (21.27)dρ dρ1 d 2 ΦΦ dφ 2 = −n2 , (21.28)where we have taken the separation constant to have the <strong>for</strong>m n 2 <strong>for</strong> laterconvenience; <strong>for</strong> the present, n is a general (complex) number.Let us first consider the case in which n ≠ 0. The second equation, (21.28), thenhas the general solutionΦ(φ) =A exp(inφ)+B exp(−inφ). (21.29)Equation (21.27), on the other h<strong>and</strong>, is the homogeneous equationρ 2 P ′′ + ρP ′ − n 2 P =0,which must be solved either by trying a power solution in ρ or by making thesubstitution ρ =expt as described in subsection 15.2.1 <strong>and</strong> so reducing it to anequation with constant coefficients. Carrying out this procedure we findP (ρ) =Cρ n + Dρ −n . (21.30)Returning to the solution (21.29) of the azimuthal equation (21.28), we cansee that if Φ, <strong>and</strong> hence u, is to be single-valued <strong>and</strong> so not change when φincreases by 2π then n must be an integer. <strong>Mathematical</strong>ly, other values of n arepermissible, but <strong>for</strong> the description of real physical situations it is clear that thislimitation must be imposed. Having thus restricted the possible values of n inone part of the solution, the same limitations must be carried over into the radialpart, (21.30). Thus we may write a particular solution of the two-dimensionalLaplace equation asu(ρ, φ) =(A cos nφ + B sin nφ)(Cρ n + Dρ −n ),726


21.3 SEPARATION OF VARIABLES IN POLAR COORDINATESwhere A, B, C, D are arbitrary constants <strong>and</strong> n is any integer.We have not yet, however, considered the solution when n =0.Inthiscase,the solutions of the separated ordinary equations (21.28) <strong>and</strong> (21.27), respectively,are easily shown to beΦ(φ) =Aφ + B,P (ρ) =C ln ρ + D.But, in order that u = P Φ is single-valued, we require A = 0, <strong>and</strong> so the solution<strong>for</strong> n = 0 is simply (absorbing B into C <strong>and</strong> D)u(ρ, φ) =C ln ρ + D.Superposing the solutions <strong>for</strong> the different allowed values of n, we can writethe general solution to Laplace’s equation in plane polars asu(ρ, φ) =(C 0 ln ρ + D 0 )+∞∑n=1(A n cos nφ + B n sin nφ)(C n ρ n + D n ρ −n ),(21.31)where n can take only integer values. Negative values of n have been omittedfrom the sum since they are already included in the terms obtained <strong>for</strong> positiven. We note that, since ln ρ is singular at ρ = 0, whenever we solve Laplace’sequation in a region containing the origin, C 0 must be identically zero.◮A circular drumskin has a supporting rim at ρ = a. If the rim is twisted so that itis displaced vertically by a small amount ɛ(sin φ +2sin2φ), whereφ is the azimuthalangle with respect to a given radius, find the resulting displacement u(ρ, φ) over the entiredrumskin.The transverse displacement of a circular drumskin is usually described by the twodimensionalwave equation. In this case, however, there is no time dependence <strong>and</strong> sou(ρ, φ) solves the two-dimensional Laplace equation, subject to the imposed boundarycondition.Referring to (21.31), since we wish to find a solution that is finite everywhere insideρ = a, werequireC 0 =0<strong>and</strong>D n =0<strong>for</strong>alln>0. Now the boundary condition at therim requiresu(a, φ) =D 0 +∞∑C n a n (A n cos nφ + B n sin nφ) =ɛ(sin φ +2sin2φ).n=1Firstly we see that we require D 0 =0<strong>and</strong>A n =0<strong>for</strong>alln. Furthermore, we musthave C 1 B 1 a = ɛ, C 2 B 2 a 2 =2ɛ <strong>and</strong> B n =0<strong>for</strong>n>2. Hence the appropriate shape <strong>for</strong> thedrumskin (valid over the whole skin, not just the rim) isu(ρ, φ) = ɛρ asin φ +2ɛρ2a 2sin 2φ = ɛρ a(sin φ + 2ρ a sin 2φ ). ◭727


PDES: SEPARATION OF VARIABLES AND OTHER METHODSLaplace’s equation in cylindrical polarsPassing to three dimensions, we now consider the solution of Laplace’s equationin cylindrical polar coordinates,(1 ∂ρ ∂u )+ 1 ∂ 2 uρ ∂ρ ∂ρ ρ 2 ∂φ 2 + ∂2 u=0. (21.32)∂z2 We note here that, even when considering a cylindrical physical system, if thereis no dependence of the physical variables on z (i.e. along the length of thecylinder) then the problem may be treated using two-dimensional plane polars,as discussed above.For the more general case, however, we proceed as previously by trying asolution of the <strong>for</strong>mu(ρ, φ, z) =P (ρ)Φ(φ)Z(z),which, on substitution into (21.32) <strong>and</strong> division through by u = P ΦZ, gives(1 dρ dP )+ 1 d 2 ΦPρdρdρ Φρ 2 dφ 2 + 1 d 2 ZZ dz 2 =0.The last term depends only on z, <strong>and</strong> the first <strong>and</strong> second (taken together) dependonly on ρ <strong>and</strong> φ. Taking the separation constant to be k 2 , we find1 d 2 ZZ dz 2 = k2 ,(1 dρ dP )+ 1 d 2 ΦPρdρdρ Φρ 2 dφ 2 + k2 =0.The first of these equations has the straight<strong>for</strong>ward solutionZ(z) =E exp(−kz)+F exp kz.Multiplying the second equation through by ρ 2 ,weobtain(ρ dρ dP )+ 1 d 2 ΦP dρ dρ Φ dφ 2 + k2 ρ 2 =0,in which the second term depends only on Φ <strong>and</strong> the other terms depend onlyon ρ. Taking the second separation constant to be m 2 , we find1 d 2 ΦΦ dφ 2 = −m2 , (21.33)ρ d (ρ dP )+(k 2 ρ 2 − m 2 )P =0. (21.34)dρ dρThe equation in the azimuthal angle φ has the very familiar solutionΦ(φ) =C cos mφ + D sin mφ.728


21.3 SEPARATION OF VARIABLES IN POLAR COORDINATESAs in the two-dimensional case, single-valuedness of u requires that m is aninteger. However, in the particular case m = 0 the solution isΦ(φ) =Cφ + D.This <strong>for</strong>m is appropriate to a solution with axial symmetry (C = 0) or one that ismultivalued, but manageably so, such as the magnetic scalar potential associatedwith a current I (in which case C = I/(2π) <strong>and</strong>D is arbitrary).Finally, the ρ-equation (21.34) may be trans<strong>for</strong>med into Bessel’s equation o<strong>for</strong>der m by writing µ = kρ. This has the solutionP (ρ) =AJ m (kρ)+BY m (kρ).The properties of these functions were investigated in chapter 16 <strong>and</strong> will notbe pursued here. We merely note that Y m (kρ) is singular at ρ = 0, <strong>and</strong> so, whenseeking solutions to Laplace’s equation in cylindrical coordinates within someregion containing the ρ = 0 axis, we require B =0.The complete separated-variable solution in cylindrical polars of Laplace’sequation ∇ 2 u = 0 is thus given byu(ρ, φ, z) =[AJ m (kρ)+BY m (kρ)][C cos mφ + D sin mφ][E exp(−kz)+F exp kz].(21.35)Of course we may use the principle of superposition to build up more generalsolutions by adding together solutions of the <strong>for</strong>m (21.35) <strong>for</strong> all allowed valuesof the separation constants k <strong>and</strong> m.◮A semi-infinite solid cylinder of radius a has its curved surface held at 0 ◦ C <strong>and</strong> its baseheld at a temperature T 0 . Find the steady-state temperature distribution in the cylinder.The physical situation is shown in figure 21.5. The steady-state temperature distributionu(ρ, φ, z) must satisfy Laplace’s equation subject to the imposed boundary conditions. Letus take the cylinder to have its base in the z = 0 plane <strong>and</strong> to extend along the positivez-axis. From (21.35), in order that u is finite everywhere in the cylinder we immediatelyrequire B =0<strong>and</strong>F = 0. Furthermore, since the boundary conditions, <strong>and</strong> hence thetemperature distribution, are axially symmetric, we require m =0,<strong>and</strong>sothegeneralsolution must be a superposition of solutions of the <strong>for</strong>m J 0 (kρ)exp(−kz) <strong>for</strong> all allowedvalues of the separation constant k.The boundary condition u(a, φ, z) = 0 restricts the allowed values of k, sincewemusthave J 0 (ka) = 0. The zeros of Bessel functions are given in most books of mathematicaltables, <strong>and</strong> we find that, to two decimal places,J 0 (x) =0 <strong>for</strong>x =2.40, 5.52, 8.65, ....Writing the allowed values of k as k n <strong>for</strong> n =1, 2, 3,... (so, <strong>for</strong> example, k 1 =2.40/a), therequired solution takes the <strong>for</strong>mu(ρ, φ, z) =∞∑A n J 0 (k n ρ)exp(−k n z).n=1729


PDES: SEPARATION OF VARIABLES AND OTHER METHODSzu =0 u =0ayxu = T 0Figure 21.5 A uni<strong>for</strong>m metal cylinder whose curved surface is kept at 0 ◦ C<strong>and</strong> whose base is held at a temperature T 0 .By imposing the remaining boundary condition u(ρ, φ, 0) = T 0 , the coefficients A n can befound in a similar way to Fourier coefficients but this time by exploiting the orthogonalityof the Bessel functions, as discussed in chapter 16. From this boundary condition werequire∞∑u(ρ, φ, 0) = A n J 0 (k n ρ)=T 0 .n=1If we multiply this expression by ρJ 0 (k r ρ) <strong>and</strong> integrate from ρ =0toρ = a, <strong>and</strong> use theorthogonality of the Bessel functions J 0 (k n ρ), then the coefficients are given by (18.91) asA n = 2T 0a 2 J1 2(k J 0 (k n ρ)ρdρ. (21.36)na) 0The integral on the RHS can be evaluated using the recurrence relation (18.92) ofchapter 16,ddz [zJ 1(z)] = zJ 0 (z),whichonsettingz = k n ρ yields1 dk n dρ [k nρJ 1 (k n ρ)] = k n ρJ 0 (k n ρ).There<strong>for</strong>e the integral in (21.36) is given by∫ a[ 1J 0 (k n ρ)ρdρ= ρJ 1 (k n ρ) = 1 aJ 1 (k n a),k n k n0∫ a730] a0


21.3 SEPARATION OF VARIABLES IN POLAR COORDINATES<strong>and</strong> the coefficients A n may be expressed asA n = 2T 0a 2 J 2 1 (k na)[ ] aJ1 (k n a)=k n2T 0k n aJ 1 (k n a) .The steady-state temperature in the cylinder is then given by∞∑ 2T 0u(ρ, φ, z) =k n aJ 1 (k n a) J 0(k n ρ)exp(−k n z). ◭n=1We note that if, in the above example, the base of the cylinder were not kept ata uni<strong>for</strong>m temperature T 0 , but instead had some fixed temperature distributionT (ρ, φ), then the solution of the problem would become more complicated. Insuch a case, the required temperature distribution u(ρ, φ, z)isingeneralnot axiallysymmetric, <strong>and</strong> so the separation constant m is not restricted to be zero but maytake any integer value. The solution will then take the <strong>for</strong>m∞∑ ∞∑u(ρ, φ, z) = J m (k nm ρ)(C nm cos mφ + D nm sin mφ) exp(−k nm z),m=0 n=1where the separation constants k nm are such that J m (k nm a) = 0, i.e. k nm a is the nthzero of the mth-order Bessel function. At the base of the cylinder we would thenrequire∞∑ ∞∑u(ρ, φ, 0) = J m (k nm ρ)(C nm cos mφ + D nm sin mφ) =T (ρ, φ).m=0 n=1(21.37)The coefficients C nm could be found by multiplying (21.37) by J q (k rq ρ)cosqφ,integrating with respect to ρ <strong>and</strong> φ over the base of the cylinder <strong>and</strong> exploitingthe orthogonality of the Bessel functions <strong>and</strong> of the trigonometric functions. TheD nm could be found in a similar way by multiplying (21.37) by J q (k rq ρ)sinqφ.Laplace’s equation in spherical polarsWe now come to an equation that is very widely applicable in physical science,namely ∇ 2 u = 0 in spherical polar coordinates:(1 ∂r 2 r 2 ∂u )+ 1 (∂∂r ∂r r 2 sin θ ∂u )1 ∂ 2 u+sin θ ∂θ ∂θ r 2 sin 2 =0. (21.38)θ ∂φ2 Our method of procedure will be as be<strong>for</strong>e; we try a solution of the <strong>for</strong>mu(r, θ, φ) =R(r)Θ(θ)Φ(φ).Substituting this in (21.38), dividing through by u = RΘΦ <strong>and</strong> multiplying by r 2 ,we obtain(1 dr 2 dR )+ 1 (dsin θ dΘ )1 d 2 Φ+R dr dr Θsinθ dθ dθ Φsin 2 =0. (21.39)θ dφ2 731


PDES: SEPARATION OF VARIABLES AND OTHER METHODSThe first term depends only on r <strong>and</strong> the second <strong>and</strong> third terms (taken together)depend only on θ <strong>and</strong> φ. Thus (21.39) is equivalent to the two equations(1 dr 2 dR )= λ, (21.40)R dr dr(1 dsin θ dΘ )1 d 2 Φ+Θsinθ dθ dθ Φsin 2 = −λ. (21.41)θ dφ2 Equation (21.40) is a homogeneous equation,r 2 d2 Rdr 2dR+2r − λR =0,drwhich can be reduced, by the substitution r =expt (<strong>and</strong> writing R(r) =S(t)), tod 2 Sdt 2This has the straight<strong>for</strong>ward solution+ dS − λS =0.dtS(t) =A exp λ 1 t + B exp λ 2 t,<strong>and</strong> so the solution to the radial equation isR(r) =Ar λ1 + Br λ2 ,where λ 1 + λ 2 = −1 <strong>and</strong>λ 1 λ 2 = −λ. We can thus take λ 1 <strong>and</strong> λ 2 as given by l<strong>and</strong> −(l +1);λ then has the <strong>for</strong>m l(l + 1). (It should be noted that at this stagenothing has been either assumed or proved about whether l is an integer.)Hence we have obtained some in<strong>for</strong>mation about the first factor in theseparated-variable solution, which will now have the <strong>for</strong>mu(r, θ, φ) = [ Ar l + Br −(l+1)] Θ(θ)Φ(φ), (21.42)where Θ <strong>and</strong> Φ must satisfy (21.41) with λ = l(l +1).The next step is to take (21.41) further. Multiplying through by sin 2 θ <strong>and</strong>substituting <strong>for</strong> λ, it too takes a separated <strong>for</strong>m:[ sin θΘddθ(sin θ dΘdθ)+ l(l +1)sin 2 θ]+ 1 d 2 Φ=0. (21.43)Φ dφ2 Taking the separation constant as m 2 , the equation in the azimuthal angle φhas the same solution as in cylindrical polars, namelyΦ(φ) =C cos mφ + D sin mφ.As be<strong>for</strong>e, single-valuedness of u requires that m is an integer; <strong>for</strong> m = 0 we againhave Φ(φ) =Cφ + D.732


21.3 SEPARATION OF VARIABLES IN POLAR COORDINATESHaving settled the <strong>for</strong>m of Φ(φ), we are left only with the equation satisfied byΘ(θ), which is(sin θ dsin θ dΘ )+ l(l +1)sin 2 θ = m 2 . (21.44)Θ dθ dθA change of independent variable from θ to µ =cosθ will reduce this to a<strong>for</strong>m <strong>for</strong> which solutions are known, <strong>and</strong> of which some study has been made inchapter 16. Puttingµ =cosθ,dµd= − sin θ,dθdθ = −(1 − µ2 ) 1/2 ddµ ,the equation <strong>for</strong> M(µ) ≡ Θ(θ) reads[d(1 − µ 2 ) dM ]]+[l(l +1)− m2dµ dµ1 − µ 2 M =0. (21.45)This equation is the associated Legendre equation, which was mentioned in subsection18.2 in the context of Sturm–Liouville equations.We recall that <strong>for</strong> the case m = 0, (21.45) reduces to Legendre’s equation, whichwas studied at length in chapter 16, <strong>and</strong> has the solutionM(µ) =EP l (µ)+FQ l (µ). (21.46)We have not solved (21.45) explicitly <strong>for</strong> general m, but the solutions were givenin subsection 18.2 <strong>and</strong> are the associated Legendre functions P m l (µ) <strong>and</strong>Qm l (µ),wherePl m (µ) =(1− µ 2 |m|/2 d|m|)dµ P l(µ), (21.47)|m|<strong>and</strong> similarly <strong>for</strong> Q m l (µ). We then haveM(µ) =EP m l (µ)+FQ m l (µ); (21.48)here m must be an integer, 0 ≤|m| ≤l. We note that if we require solutions toLaplace’s equation that are finite when µ =cosθ = ±1 (i.e. on the polar axiswhere θ =0,π), then we must have F = 0 in (21.46) <strong>and</strong> (21.48) since Q m l (µ)diverges at µ = ±1.It will be remembered that one of the important conditions <strong>for</strong> obtainingfinite polynomial solutions of Legendre’s equation is that l is an integer ≥ 0.This condition there<strong>for</strong>e applies also to the solutions (21.46) <strong>and</strong> (21.48) <strong>and</strong> isreflected back into the radial part of the general solution given in (21.42).Now that the solutions of each of the three ordinary differential equationsgoverning R, Θ <strong>and</strong> Φ have been obtained, we may assemble a complete separated-733


PDES: SEPARATION OF VARIABLES AND OTHER METHODSvariable solution of Laplace’s equation in spherical polars. It isu(r, θ, φ) =(Ar l + Br −(l+1) )(C cos mφ + D sin mφ)[EP m l (cos θ)+FQ m l (cos θ)],(21.49)where the three bracketted factors are connected only through the integer parametersl <strong>and</strong> m, 0≤|m| ≤l. As be<strong>for</strong>e, a general solution may be obtainedby superposing solutions of this <strong>for</strong>m <strong>for</strong> the allowed values of the separationconstants l <strong>and</strong> m. As mentioned above, if the solution is required to be finite onthe polar axis then F = 0 <strong>for</strong> all l <strong>and</strong> m.◮An uncharged conducting sphere of radius a is placed at the origin in an initially uni<strong>for</strong>melectrostatic field E. Show that it behaves as an electric dipole.The uni<strong>for</strong>m field, taken in the direction of the polar axis, has an electrostatic potentialu = −Ez = −Ercos θ,where u is arbitrarily taken as zero at z = 0. This satisfies Laplace’s equation ∇ 2 u =0,asmust the potential v when the sphere is present; <strong>for</strong> large r the asymptotic <strong>for</strong>m of v muststill be −Er cos θ.Since the problem is clearly axially symmetric, we have immediately that m =0,<strong>and</strong>sincewerequirev to be finite on the polar axis we must have F = 0 in (21.49). There<strong>for</strong>ethe solution must be of the <strong>for</strong>m∞∑v(r, θ, φ) = (A l r l + B l r −(l+1) )P l (cos θ).l=0Now the cos θ-dependence of v <strong>for</strong> large r indicates that the (θ, φ)-dependence of v(r, θ, φ)is given by P1 0 (cos θ) =cosθ. Thus the r-dependence of v must also correspond to anl = 1 solution, <strong>and</strong> the most general such solution (outside the sphere, i.e. <strong>for</strong> r ≥ a) isv(r, θ, φ) =(A 1 r + B 1 r −2 )P 1 (cos θ).Theasymptotic<strong>for</strong>mofv <strong>for</strong> large r immediately gives A 1 = −E <strong>and</strong> so yields the solution(v(r, θ, φ) = −Er + B )1cos θ.r 2Since the sphere is conducting, it is an equipotential region <strong>and</strong> so v must not depend onθ <strong>for</strong> r = a. This can only be the case if B 1 /a 2 = Ea, thus fixing B 1 . The final solution isthere<strong>for</strong>e)v(r, θ, φ) =−Er(1 − a3cos θ.r 3Since a dipole of moment p gives rise to a potential p/(4πɛ 0 r 2 ), this result shows that thesphere behaves as a dipole of moment 4πɛ 0 a 3 E, because of the charge distribution inducedon its surface; see figure 21.6. ◭Often the boundary conditions are not so easily met, <strong>and</strong> it is necessary touse the mutual orthogonality of the associated Legendre functions (<strong>and</strong> thetrigonometric functions) to obtain the coefficients in the general solution.734


21.3 SEPARATION OF VARIABLES IN POLAR COORDINATES− +++−− − ++− − θ +− a +− +−+−+− + − +Figure 21.6 Induced charge <strong>and</strong> field lines associated with a conductingsphere placed in an initially uni<strong>for</strong>m electrostatic field.◮A hollow split conducting sphere of radius a is placed at the origin. If one half of itssurface is charged to a potential v 0 <strong>and</strong> the other half is kept at zero potential, find thepotential v inside <strong>and</strong> outside the sphere.Let us choose the top hemisphere to be charged to v 0 <strong>and</strong> the bottom hemisphere to beat zero potential, with the plane in which the two hemispheres meet perpendicular to thepolar axis; this is shown in figure 21.7. The boundary condition then becomes{v0 <strong>for</strong> 0


PDES: SEPARATION OF VARIABLES AND OTHER METHODSzav = v 0θrφyxv =0−aFigure 21.7 A hollow split conducting sphere with its top half charged to apotential v 0 <strong>and</strong> its bottom half at zero potential.where in the last line we have used (21.50). The integrals of the Legendre polynomials areeasily evaluated (see exercise 17.3) <strong>and</strong> we findA 0 = v 02 , A 1 = 3v 04a , A 2 =0, A 3 = − 7v 016a 3 , ··· ,so that the required solution inside the sphere isv(r, θ, φ) = v 02[1+ 3r2a P 1(cos θ) − 7r38a 3 P 3(cos θ)+···Outside the sphere (<strong>for</strong> r>a) we require the solution to be bounded as r tends toinfinity <strong>and</strong> so in (21.51) we must have A l =0<strong>for</strong>alll. In this case, by imposing theboundary condition at r = a we require∞∑v(a, θ, φ) = B l a −(l+1) P l (cos θ),l=0where v(a, θ, φ) is given by (21.50). Following the above argument the coefficients in theexpansion are given byB l a −(l+1) =2l +12so that the required solution outside the sphere isv(r, θ, φ) = v 0a2r∫ 1v 0 P l (µ)dµ,[1+ 3a2r P 1(cos θ) − 7a38r 3 P 3(cos θ)+···0].]. ◭736


21.3 SEPARATION OF VARIABLES IN POLAR COORDINATESIn the above example, on the equator of the sphere (i.e. at r = a <strong>and</strong> θ = π/2)the potential is given byv(a, π/2,φ)=v 0 /2,i.e. mid-way between the potentials of the top <strong>and</strong> bottom hemispheres. This isso because a Legendre polynomial expansion of a function behaves in the sameway as a Fourier series expansion, in that it converges to the average of the twovalues at any discontinuities present in the original function.If the potential on the surface of the sphere had been given as a function of θ<strong>and</strong> φ, then we would have had to consider a double series summed over l <strong>and</strong>m (<strong>for</strong> −l ≤ m ≤ l), since, in general, the solution would not have been axiallysymmetric.Finally, we note in general that, when obtaining solutions of Laplace’s equationin spherical polar coordinates, one finds that, <strong>for</strong> solutions that are finite on thepolar axis, the angular part of the solution is given byΘ(θ)Φ(φ) =P m l (cos θ)(C cos mφ + D sin mφ),where l <strong>and</strong> m are integers with −l ≤ m ≤ l. This general <strong>for</strong>m is sufficientlycommon that particular functions of θ <strong>and</strong> φ called spherical harmonics aredefined <strong>and</strong> tabulated (see section 18.3).21.3.2 Other equations in polar coordinatesThe development of the solutions of ∇ 2 u = 0 carried out in the previous subsectioncan be employed to solve other equations in which the ∇ 2 operator appears. Sincewe have discussed the general method in some depth already, only an outline ofthe solutions will be given here.Let us first consider the wave equation∇ 2 u = 1 ∂ 2 uc 2 ∂t 2 , (21.52)<strong>and</strong> look <strong>for</strong> a separated solution of the <strong>for</strong>m u = F(r)T (t), so that initially weare separating only the spatial <strong>and</strong> time dependences. Substituting this <strong>for</strong>m into(21.52) <strong>and</strong> taking the separation constant as k 2 we obtain∇ 2 F + k 2 d 2 TF =0,dt 2 + k2 c 2 T =0. (21.53)The second equation has the simple solutionT (t) =A exp(iωt)+B exp(−iωt), (21.54)where ω = kc; this may also be expressed in terms of sines <strong>and</strong> cosines, of course.The first equation in (21.53) is referred to as Helmholtz’s equation; we discuss itbelow.737


PDES: SEPARATION OF VARIABLES AND OTHER METHODSWe may treat the diffusion equationκ∇ 2 u = ∂u∂tin a similar way. Separating the spatial <strong>and</strong> time dependences by assuming asolution of the <strong>for</strong>m u = F(r)T (t), <strong>and</strong> taking the separation constant as k 2 ,wefind∇ 2 F + k 2 dTF =0,dt + k2 κT =0.Just as in the case of the wave equation, the spatial part of the solution satisfiesHelmholtz’s equation. It only remains to consider the time dependence, whichhas the simple solutionT (t) =A exp(−k 2 κt).Helmholtz’s equation is clearly of central importance in the solutions of thewave <strong>and</strong> diffusion equations. It can be solved in polar coordinates in much thesame way as Laplace’s equation, <strong>and</strong> indeed reduces to Laplace’s equation whenk = 0. There<strong>for</strong>e, we will merely sketch the method of its solution in each of thethree polar coordinate systems.Helmholtz’s equation in plane polarsIn two-dimensional plane polar coordinates, Helmholtz’s equation takes the <strong>for</strong>m(1 ∂ρ ∂F )+ 1 ∂ 2 Fρ ∂ρ ∂ρ ρ 2 ∂φ 2 + k2 F =0.If we try a separated solution of the <strong>for</strong>m F(r) = P (ρ)Φ(φ), <strong>and</strong> take theseparation constant as m 2 , we findd 2 Φdφ 2 + m2 φ =0,d 2 Pdρ 2 + 1 ( )dPρ dρ + k 2 − m2ρ 2 P =0.As <strong>for</strong> Laplace’s equation, the angular part has the familiar solution (if m ≠0)Φ(φ) =A cos mφ + B sin mφ,or an equivalent <strong>for</strong>m in terms of complex exponentials. The radial equationdiffers from that found in the solution of Laplace’s equation, but by making thesubstitution µ = kρ it is easily trans<strong>for</strong>med into Bessel’s equation of order m(discussed in chapter 16), <strong>and</strong> has the solutionP (ρ) =CJ m (kρ)+DY m (kρ),where Y m is a Bessel function of the second kind, which is infinite at the origin738


21.3 SEPARATION OF VARIABLES IN POLAR COORDINATES<strong>and</strong> is not to be confused with a spherical harmonic (these are written with asuperscript as well as a subscript).Putting the two parts of the solution together we haveF(ρ, φ) =[A cos mφ + B sin mφ][CJ m (kρ)+DY m (kρ)]. (21.55)Clearly, <strong>for</strong> solutions of Helmholtz’s equation that are required to be finite at theorigin, we must set D =0.◮Find the four lowest frequency modes of oscillation of a circular drumskin of radius awhose circumference is held fixed in a plane.The transverse displacement u(r,t) of the drumskin satisfies the two-dimensional waveequation∇ 2 u = 1 ∂ 2 uc 2 ∂t , 2with c 2 = T/σ,whereT is the tension of the drumskin <strong>and</strong> σ is its mass per unit area.From (21.54) <strong>and</strong> (21.55) a separated solution of this equation, in plane polar coordinates,that is finite at the origin isu(ρ, φ, t) =J m (kρ)(A cos mφ + B sin mφ)exp(±iωt),where ω = kc. Since we require the solution to be single-valued we must have m as aninteger. Furthermore, if the drumskin is clamped at its outer edge ρ = a then we alsorequire u(a, φ, t) = 0. Thus we needJ m (ka) =0,which in turn restricts the allowed values of k. The zeros of Bessel functions can beobtained from most books of tables; the first few areJ 0 (x) =0 <strong>for</strong>x ≈ 2.40, 5.52, 8.65,...,J 1 (x) =0 <strong>for</strong>x ≈ 3.83, 7.02, 10.17,...,J 2 (x) =0 <strong>for</strong>x ≈ 5.14, 8.42, 11.62 ... .The smallest value of x <strong>for</strong> which any of the Bessel functions is zero is x ≈ 2.40, whichoccurs <strong>for</strong> J 0 (x). Thus the lowest-frequency mode has k =2.40/a <strong>and</strong> angular frequencyω =2.40c/a. Sincem = 0 <strong>for</strong> this mode, the shape of the drumskin is(u ∝ J 0 2.40 ρ );athis is illustrated in figure 21.8.Continuing in the same way, the next three modes are given byω =3.83 c (a , u ∝ J 1 3.83 ρ ) (cos φ, J 1 3.83 ρ )sin φ;aaω =5.14 c (a , u ∝ J 2 5.14 ρ ) (cos 2φ, J 2 5.14 ρ )sin 2φ;aaω =5.52 c (a , u ∝ J 0 5.52 ρ ).aThese modes are also shown in figure 21.8. We note that the second <strong>and</strong> third frequencieshave two corresponding modes of oscillation; these frequencies are there<strong>for</strong>e two-folddegenerate. ◭739


PDES: SEPARATION OF VARIABLES AND OTHER METHODSaω =2.40c/aω =3.83c/aω =5.14c/aω =5.52c/aFigure 21.8 The modes of oscillation with the four lowest frequencies <strong>for</strong> acircular drumskin of radius a. The dashed lines indicate the nodes, where thedisplacement of the drumskin is always zero.Helmholtz’s equation in cylindrical polarsGeneralising the above method to three-dimensional cylindrical polars is straight<strong>for</strong>ward,<strong>and</strong> following a similar procedure to that used <strong>for</strong> Laplace’s equationwe find the separated solution of Helmholtz’s equation takes the <strong>for</strong>m[ (√ ) (√ )]F(ρ, φ, z) = AJ m k2 − α 2 ρ + BY m k2 − α 2 ρ× (C cos mφ + D sin mφ)[E exp(iαz)+F exp(−iαz)],where α <strong>and</strong> m are separation constants. We note that the angular part of thesolution is the same as <strong>for</strong> Laplace’s equation in cylindrical polars.Helmholtz’s equation in spherical polarsIn spherical polars, we find again that the angular parts of the solution Θ(θ)Φ(φ)are identical to those of Laplace’s equation in this coordinate system, i.e. they arethe spherical harmonics Yl m (θ, φ), <strong>and</strong> so we shall not discuss them further.The radial equation in this case is given byr 2 R ′′ +2rR ′ +[k 2 r 2 − l(l +1)]R =0, (21.56)which has an additional term k 2 r 2 R compared with the radial equation <strong>for</strong> theLaplace solution. The equation (21.56) looks very much like Bessel’s equation.In fact, by writing R(r) =r −1/2 S(r) <strong>and</strong> making the change of variable µ = kr,it can be reduced to Bessel’s equation of order l + 1 2, which has as its solutionsS(µ) = J l+1/2 (µ) <strong>and</strong> Y l+1/2 (µ) (see section 18.6). The separated solution to740


21.3 SEPARATION OF VARIABLES IN POLAR COORDINATESHelmholtz’s equation in spherical polars is thusF(r, θ, φ) =r −1/2 [AJ l+1/2 (kr)+BY l+1/2 (kr)](C cos mφ + D sin mφ)×[EPl m (cos θ)+FQ m l (cos θ)]. (21.57)For solutions that are finite at the origin we require B = 0, <strong>and</strong> <strong>for</strong> solutionsthat are finite on the polar axis we require F = 0. It is worth mentioning thatthe solutions proportional to r −1/2 J l+1/2 (kr) <strong>and</strong>r −1/2 Y l+1/2 (kr), when suitablynormalised, are called spherical Bessel functions of the first <strong>and</strong> second kind,respectively, <strong>and</strong> are denoted by j l (kr) <strong>and</strong>n l (µ) (see section 18.6).As mentioned at the beginning of this subsection, the separated solution ofthe wave equation in spherical polars is the product of a time-dependent part(21.54) <strong>and</strong> a spatial part (21.57). It will be noticed that, although this solutioncorresponds to a solution of definite frequency ω = kc, the zeros of the radialfunction j l (kr) are not equally spaced in r, except <strong>for</strong> the case l = 0 involvingj 0 (kr), <strong>and</strong> so there is no precise wavelength associated with the solution.To conclude this subsection, let us mention briefly the Schrödinger equation<strong>for</strong> the electron in a hydrogen atom, the nucleus of which is taken at the origin<strong>and</strong> is assumed massive compared with the electron. Under these circumstancesthe Schrödinger equation is− 22m ∇2 u −e2 u4πɛ 0 r = i∂u ∂t .For a ‘stationary-state’ solution, <strong>for</strong> which the energy is a constant E <strong>and</strong> the timedependentfactor T in u is given by T (t) =A exp(−iEt/), the above equation issimilar to, but not quite the same as, the Helmholtz equation. § However, as withthe wave equation, the angular parts of the solution are identical to those <strong>for</strong>Laplace’s equation <strong>and</strong> are expressed in terms of spherical harmonics.The important point to note is that <strong>for</strong> any equation involving ∇ 2 , provided θ<strong>and</strong> φ do not appear in the equation other than as part of ∇ 2 , a separated-variablesolution in spherical polars will always lead to spherical harmonic solutions. Thisis the case <strong>for</strong> the Schrödinger equation describing an atomic electron in a centralpotential V (r).21.3.3 Solution by expansionIt is sometimes possible to use the uniqueness theorem discussed in the previouschapter, together with the results of the last few subsections, in which Laplace’sequation (<strong>and</strong> other equations) were considered in polar coordinates, to obtainsolutions of such equations appropriate to particular physical situations.§ For the solution by series of the r-equation in this case the reader may consult, <strong>for</strong> example, L.Schiff, Quantum Mechanics (New York: McGraw-Hill, 1955), p. 82.741


PDES: SEPARATION OF VARIABLES AND OTHER METHODSPzθr−aOayxFigure 21.9 The polar axis Oz is taken as normal to the plane of the ring ofmatter <strong>and</strong> passing through its centre.We will illustrate the method <strong>for</strong> Laplace’s equation in spherical polars <strong>and</strong> firstassume that the required solution of ∇ 2 u = 0 can be written as a superpositionin the normal way:u(r, θ, φ) =∞∑l=0 m=−ll∑(Ar l + Br −(l+1) )Pl m (cos θ)(C cos mφ + D sin mφ).(21.58)Here, all the constants A, B, C, D may depend upon l <strong>and</strong> m, <strong>and</strong> we haveassumed that the required solution is finite on the polar axis. As usual, boundaryconditions of a physical nature will then fix or eliminate some of the constants;<strong>for</strong> example, u finite at the origin implies all B = 0, or axial symmetry impliesthat only m = 0 terms are present.The essence of the method is then to find the remaining constants by determiningu at values of r, θ, φ <strong>for</strong> which it can be evaluated by other means, e.g. by directcalculation on an axis of symmetry. Once the remaining constants have been fixedby these special considerations to have particular values, the uniqueness theoremcan be invoked to establish that they must have these values in general.◮Calculate the gravitational potential at a general point in space due to a uni<strong>for</strong>m ring ofmatter of radius a <strong>and</strong> total mass M.Everywhere except on the ring the potential u(r) satisfies the Laplace equation, <strong>and</strong> so ifwe use polar coordinates with the normal to the ring as polar axis, as in figure 21.9, asolution of the <strong>for</strong>m (21.58) can be assumed.We expect the potential u(r, θ, φ) totendtozeroasr →∞, <strong>and</strong> also to be finite at r =0.At first sight this might seem to imply that all A <strong>and</strong> B, <strong>and</strong> hence u, must be identicallyzero, an unacceptable result. In fact, what it means is that different expressions must applyto different regions of space. On the ring itself we no longer have ∇ 2 u = 0 <strong>and</strong> so it is not742


21.3 SEPARATION OF VARIABLES IN POLAR COORDINATESsurprising that the <strong>for</strong>m of the expression <strong>for</strong> u changes there. Let us there<strong>for</strong>e take twoseparate regions.In the region r>a(i) we must have u → 0asr →∞, implying that all A =0,<strong>and</strong>(ii) the system is axially symmetric <strong>and</strong> so only m = 0 terms appear.With these restrictions we can write as a trial <strong>for</strong>m∞∑u(r, θ, φ) = B l r −(l+1) Pl 0 (cos θ). (21.59)l=0The constants B l are still to be determined; this we do by calculating directly the potentialwhere this can be done simply – in this case, on the polar axis.Considering a point P on the polar axis at a distance z (> a) from the plane of the ring(taken as θ = π/2), all parts of the ring are at a distance (z 2 + a 2 ) 1/2 from it. The potentialat P is thus straight<strong>for</strong>wardlyGMu(z,0,φ)=− , (21.60)(z 2 + a 2 )1/2where G is the gravitational constant. This must be the same as (21.59) <strong>for</strong> the particularvalues r = z, θ =0,<strong>and</strong>φ undefined. Since Pl 0(cos θ) =P l(cos θ) withP l (1) = 1, puttingr = z in (21.59) gives∞∑ B lu(z,0,φ)= . (21.61)zl+1 l=0However, exp<strong>and</strong>ing (21.60) <strong>for</strong> z>a(as it applies to this region of space) we obtainu(z,0,φ)=− GM [1 − 1 ( a) 2 3( a) 4+ − ···],z 2 z 8 zwhich on comparison with (21.61) gives §B 0 = −GM,B 2l = − GMa2l (−1) l (2l − 1)!!<strong>for</strong> l ≥ 1, (21.62)2 l l!B 2l+1 =0.We now conclude the argument by saying that if a solution <strong>for</strong> a general point (r, θ, φ)exists at all, which of course we very much expect on physical grounds, then it must be(21.59) with the B l given by (21.62). This is so because thus defined it is a function withno arbitrary constants <strong>and</strong> which satisfies all the boundary conditions, <strong>and</strong> the uniquenesstheorem states that there is only one such function. The expression <strong>for</strong> the potential in theregion r>ais there<strong>for</strong>e[]u(r, θ, φ) =− GM ∞∑ (−1) l (2l − 1)!!( a) 2l1+P2l (cos θ) .r2 l l! rl=1The expression <strong>for</strong> r


PDES: SEPARATION OF VARIABLES AND OTHER METHODSComparing this expression <strong>for</strong> r = z, θ = 0 with the z


21.3 SEPARATION OF VARIABLES IN POLAR COORDINATESwhere the coefficients R l (r) <strong>and</strong>F l (r) in the Legendre polynomial expansionsare functions of r. Since in any particular problem ρ is given, we can find thecoefficients F l (r) in the expansion in the usual way (see subsection 18.1.2). It thenonly remains to find the coefficients R l (r) in the expansion of the solution u.Writing ∇ 2 in spherical polars <strong>and</strong> substituting (21.64) <strong>and</strong> (21.65) into (21.63)we obtain∞∑[Pl (cos θ)l=0r 2ddr(r 2 dR )l+ R ldrdr 2 sin θ dθ(sin θ dP )]l(cos θ)=dθ∞∑F l (r)P l (cos θ).l=0(21.66)However, if, in equation (21.44) of our discussion of the angular part of thesolution to Laplace’s equation, we set m = 0 we conclude that(1 dsin θ dP )l(cos θ)= −l(l +1)P l (cos θ).sin θ dθ dθSubstituting this into (21.66), we find that the LHS is greatly simplified <strong>and</strong> weobtain∞∑[ ( 1 dr 2 r 2 dR )l− l(l +1)R ]∞∑ldr dr r 2 P l (cos θ) = F l (r)P l (cos θ).l=0This relation is most easily satisfied by equating terms on both sides <strong>for</strong> eachvalue of l separately, so that <strong>for</strong> l =0, 1, 2,... we have(1 dr 2 r 2 dR )l− l(l +1)R ldr dr r 2 = F l (r). (21.67)This is an ODE in which F l (r) is given, <strong>and</strong> it can there<strong>for</strong>e be solved <strong>for</strong>R l (r). The solution to Poisson’s equation, u, is then obtained by making thesuperposition (21.64).◮In a certain system, the electric charge density ρ is distributed as follows:{Ar cos θ <strong>for</strong> 0 ≤ r


PDES: SEPARATION OF VARIABLES AND OTHER METHODSThis can be rearranged to giver 2 R 1 ′′ +2rR 1 ′ − 2R 1 = − Ar3 ,ɛ 0where the prime denotes differentiation with respect to r. The LHS is homogeneous <strong>and</strong>the equation can be reduced by the substitution r =expt, <strong>and</strong> writing R 1 (r) =S(t), to¨S + Ṡ − 2S = − A ɛ 0exp 3t, (21.68)where the dots indicate differentiation with respect to t.This is an inhomogeneous second-order ODE with constant coefficients <strong>and</strong> can bestraight<strong>for</strong>wardly solved by the methods of subsection 15.2.1 to giveRecalling that r =expt we findS(t) =c 1 exp t + c 2 exp(−2t) −A10ɛ 0exp 3t.R 1 (r) =c 1 r + c 2 r −2 −A r 3 .10ɛ 0Since we are interested in the region r


21.4 INTEGRAL TRANSFORM METHODS21.4 Integral trans<strong>for</strong>m methodsIn the method of separation of variables our aim was to keep the independentvariables in a PDE as separate as possible. We now discuss the use of integraltrans<strong>for</strong>ms in solving PDEs, a method by which one of the independent variablescan be eliminated from the differential coefficients. It will be assumed that thereader is familiar with Laplace <strong>and</strong> Fourier trans<strong>for</strong>ms <strong>and</strong> their properties, asdiscussed in chapter 13.The method consists simply of trans<strong>for</strong>ming the PDE into one containingderivatives with respect to a smaller number of variables. Thus, if the originalequation has just two independent variables, it may be possible to reduce thePDE into a soluble ODE. The solution obtained can then (where possible) betrans<strong>for</strong>med back to give the solution of the original PDE. As we shall see,boundary conditions can usually be incorporated in a natural way.Which sort of trans<strong>for</strong>m to use, <strong>and</strong> the choice of the variable(s) with respectto which the trans<strong>for</strong>m is to be taken, is a matter of experience; we illustrate thisin the example below. In practice, trans<strong>for</strong>ms can be taken with respect to eachvariable in turn, <strong>and</strong> the trans<strong>for</strong>mation that af<strong>for</strong>ds the greatest simplificationcan be pursued further.◮A semi-infinite tube of constant cross-section contains initially pure water. At time t =0,one end of the tube is put into contact with a salt solution <strong>and</strong> maintained at a concentrationu 0 . Find the total amount of salt that has diffused into the tube after time t, if the diffusionconstant is κ.The concentration u(x, t) attimet <strong>and</strong> distance x from the end of the tube satisfies thediffusion equationκ ∂2 u∂x = ∂u2 ∂t , (21.71)which has to be solved subject to the boundary conditions u(0,t)=u 0 <strong>for</strong> all t <strong>and</strong>u(x, 0) = 0 <strong>for</strong> all x>0.Since we are interested only in t>0, the use of the Laplace trans<strong>for</strong>m is suggested.Furthermore, it will be recalled from chapter 13 that one of the major virtues of Laplacetrans<strong>for</strong>mations is the possibility they af<strong>for</strong>d of replacing derivatives of functions by simplemultiplication by a scalar. If the derivative with respect to time were so removed, equation(21.71) would contain only differentiation with respect to a single variable. Let us there<strong>for</strong>etake the Laplace trans<strong>for</strong>m of (21.71) with respect to t:∫ ∞∫κ ∂2 u∞0 ∂x exp(−st) dt = ∂uexp(−st) dt.2 0 ∂tOn the LHS the (double) differentiation is with respect to x, whereas the integration iswith respect to the independent variable t. There<strong>for</strong>e the derivative can be taken outsidethe integral. Denoting the Laplace trans<strong>for</strong>m of u(x, t) byū(x, s) <strong>and</strong> using result (13.57)to rewrite the trans<strong>for</strong>m of the derivative on the RHS (or by integrating directly by parts),we obtainκ ∂2 ū= sū(x, s) − u(x, 0).∂x2 But from the boundary condition u(x, 0) = 0 the last term on the RHS vanishes, <strong>and</strong> the747


PDES: SEPARATION OF VARIABLES AND OTHER METHODSsolution is immediate:ū(x, s) =A exp(√ ) ( √ )s sκ x + B exp −κ x ,where the constants A <strong>and</strong> B may depend on s.We require u(x, t) → 0asx →∞<strong>and</strong> so we must also have ū(∞,s) = 0; consequentlywe require that A = 0. The value of B is determined by the need <strong>for</strong> u(0,t)=u 0 <strong>and</strong> hencethat∫ ∞ū(0,s)= u 0 exp(−st) dt = u 00s .We thus conclude that the appropriate expression <strong>for</strong> the Laplace trans<strong>for</strong>m of u(x, t) isū(x, s) = u ( √ )0 ss exp −κ x . (21.72)To obtain u(x, t) from this result requires the inversion of this trans<strong>for</strong>m – a task that isgenerally difficult <strong>and</strong> requires a contour integration. This is discussed in chapter 24, but<strong>for</strong> completeness we note that the solution is( )] xu(x, t) =u 0[1 − erf √ ,4κtwhere erf(x) is the error function discussed in the Appendix. (The more complete sets ofmathematical tables list this inverse Laplace trans<strong>for</strong>m.)In the present problem, however, an alternative method is available. Let w(t) betheamount of salt that has diffused into the tube in time t; then<strong>and</strong> its trans<strong>for</strong>m is given by¯w(s) ===w(t) =∫ ∞0∫ ∞0∫ ∞0∫ ∞0dt exp(−st)dx∫ ∞0ū(x, s) dx.u(x, t) dx,∫ ∞0u(x, t) dxu(x, t)exp(−st) dtSubstituting <strong>for</strong> ū(x, s) from (21.72) into the last integral <strong>and</strong> integrating, we obtain¯w(s) =u 0 κ 1/2 s −3/2 .This expression is much simpler to invert, <strong>and</strong> referring to the table of st<strong>and</strong>ard Laplacetrans<strong>for</strong>ms (table 13.1) we findw(t) =2(κ/π) 1/2 u 0 t 1/2 ,which is thus the required expression <strong>for</strong> the amount of diffused salt at time t. ◭The above example shows that in some circumstances the use of a Laplacetrans<strong>for</strong>mation can greatly simplify the solution of a PDE. However, it will havebeen observed that (as with ODEs) the easy elimination of some derivatives isusually paid <strong>for</strong> by the introduction of a difficult inverse trans<strong>for</strong>mation. Thisproblem, although still present, is less severe <strong>for</strong> Fourier trans<strong>for</strong>mations.748


21.4 INTEGRAL TRANSFORM METHODS◮An infinite metal bar has an initial temperature distribution f(x) along its length. Findthe temperature distribution at a later time t.We are interested in values of x from −∞ to ∞, which suggests Fourier trans<strong>for</strong>mationwith respect to x. Assuming that the solution obeys the boundary conditions u(x, t) → 0<strong>and</strong> ∂u/∂x → 0as|x| →∞, we may Fourier-trans<strong>for</strong>m the one-dimensional diffusionequation (21.71) to obtain∫κ ∞√2π−∞∂ 2 u(x, t)∂x 2 exp(−ikx) dx = 1 √2π∂∂t∫ ∞−∞u(x, t)exp(−ikx) dx,where on the RHS we have taken the partial derivative with respect to t outside theintegral. Denoting the Fourier trans<strong>for</strong>m of u(x, t) byũ(k, t), <strong>and</strong> using equation (13.28) torewrite the Fourier trans<strong>for</strong>m of the second derivative on the LHS, we then have−κk 2 ∂ũ(k, t)ũ(k, t) = .∂tThis first-order equation has the simple solutionũ(k, t) =ũ(k, 0) exp(−κk 2 t),where the initial conditions giveũ(k, 0) = √ 1 ∫ ∞u(x, 0) exp(−ikx) dx2π −∞= √ 1 ∫ ∞f(x)exp(−ikx) dx = ˜f(k).2πThus we may write the Fourier trans<strong>for</strong>m of the solution as−∞ũ(k, t) =˜f(k)exp(−κk 2 t)= √ 2π ˜f(k)˜G(k, t), (21.73)where we have defined the function ˜G(k, t) =( √ 2π) −1 exp(−κk 2 t). Since ũ(k, t) canbewritten as the product of two Fourier trans<strong>for</strong>ms, we can use the convolution theorem,subsection 13.1.7, to write the solution asu(x, t) =∫ ∞−∞G(x − x ′ ,t)f(x ′ ) dx ′ ,where G(x, t) is the Green’s function <strong>for</strong> this problem (see subsection 15.2.5). This functionis the inverse Fourier trans<strong>for</strong>m of ˜G(k, t) <strong>and</strong> is thus given byG(x, t) = 12π= 12π∫ ∞−∞∫ ∞−∞Completing the square in the integr<strong>and</strong> we findG(x, t) = 1 ( ) ∫ ∞2π exp − x24κt= 1 (2π exp − x24κt= 1 √4πκtexpexp(−κk 2 t)exp(ikx) dk[exp −κt) ∫ ∞(− x24κtexp−∞exp−∞),(k 2 − ixκt k )]dk.[−κt(k − ix ) ] 2dk2κt(−κtk ′2) dk ′where in the second line we have made the substitution k ′ = k − ix/(2κt), <strong>and</strong> in the last749


PDES: SEPARATION OF VARIABLES AND OTHER METHODSut 1t 2t 3x = axFigure 21.10 Diffusion of heat from a point source in a metal bar: the curvesshow the temperature u at position x <strong>for</strong> various times t 1


21.5 INHOMOGENEOUS PROBLEMS – GREEN’S FUNCTIONSthe infinite range in x, rather than of the trans<strong>for</strong>m method itself. In fact themethod of separation of variables would yield the same solutions, since in theinfinite-range case the separation constant is not restricted to take on an infiniteset of discrete values but may have any real value, with the result that the sumover λ becomes an integral, as mentioned at the end of section 21.2.◮An infinite metal bar has an initial temperature distribution f(x) along its length. Findthe temperature distribution at a later time t using the method of separation of variables.This is the same problem as in the previous example, but we now seek a solution byseparating variables. From (21.12) a separated solution <strong>for</strong> the one-dimensional diffusionequation is given byu(x, t) =[A exp(iλx)+B exp(−iλx)] exp(−κλ 2 t),where −λ 2 is the separation constant. Since the bar is infinite we do not require thesolution to take a given <strong>for</strong>m at any finite value of x (<strong>for</strong> instance at x =0)<strong>and</strong>sothereis no restriction on λ other than its being real. There<strong>for</strong>e instead of the superposition ofsuch solutions in the <strong>for</strong>m of a sum over allowed values of λ we have an integral overall λ,u(x, t) = √ 1 ∫ ∞A(λ)exp(−κλ 2 t)exp(iλx) dλ, (21.75)2π−∞where in taking λ from −∞ to ∞ we need include only one of the complex exponentials;we have taken a factor 1/ √ 2π out of A(λ) <strong>for</strong> convenience. We can see from (21.75)that the expression <strong>for</strong> u(x, t) has the <strong>for</strong>m of an inverse Fourier trans<strong>for</strong>m (where λ isthe trans<strong>for</strong>m variable). There<strong>for</strong>e, Fourier-trans<strong>for</strong>ming both sides <strong>and</strong> using the Fourierinversion theorem, we findũ(λ, t) =A(λ)exp(−κλ 2 t).Now, the initial boundary condition requiresu(x, 0) = √ 1 ∫ ∞A(λ)exp(iλx) dλ = f(x),2π−∞from which, using the Fourier inversion theorem once more, we see that A(λ) =˜f(λ).There<strong>for</strong>e we haveũ(λ, t) =˜f(λ)exp(−κλ 2 t),which is identical to (21.73) in the previous example (but with k replaced by λ), <strong>and</strong> henceleads to the same result. ◭21.5 Inhomogeneous problems – Green’s functionsIn chapters 15 <strong>and</strong> 17 we encountered Green’s functions <strong>and</strong> found them a usefultool <strong>for</strong> solving inhomogeneous linear ODEs. We now discuss their usefulness insolving inhomogeneous linear PDEs.For the sake of brevity we shall again denote a linear PDE byLu(r) =ρ(r), (21.76)where L is a linear partial differential operator. For example, in Laplace’s equation751


PDES: SEPARATION OF VARIABLES AND OTHER METHODSwe have L = ∇ 2 , whereas <strong>for</strong> Helmholtz’s equation L = ∇ 2 +k 2 . Note that we havenot specified the dimensionality of the problem, <strong>and</strong> (21.76) may, <strong>for</strong> example,represent Poisson’s equation in two or three (or more) dimensions. The readerwill also notice that <strong>for</strong> the sake of simplicity we have not included any timedependence in (21.76). Nevertheless, the following discussion can be generalisedto include it.As we discussed in subsection 20.3.2, a problem is inhomogeneous if the factthat u(r) is a solution does not imply that any constant multiple λu(r) isalsoasolution. This inhomogeneity may derive from either the PDE itself or from theboundary conditions imposed on the solution.In our discussion of Green’s function solutions of inhomogeneous ODEs (seesubsection 15.2.5) we dealt with inhomogeneous boundary conditions by making asuitable change of variable such that in the new variable the boundary conditionswere homogeneous. In an analogous way, as illustrated in the final exampleof section 21.2, it is usually possible to make a change of variables in PDEs totrans<strong>for</strong>m between inhomogeneity of the boundary conditions <strong>and</strong> inhomogeneityof the equation. There<strong>for</strong>e let us assume <strong>for</strong> the moment that the boundaryconditions imposed on the solution u(r) of (21.76) are homogeneous. This mostcommonly means that if we seek a solution to (21.76) in some region V thenon the surface S that bounds V the solution obeys the conditions u(r) =0or∂u/∂n =0,where∂u/∂n is the normal derivative of u at the surface S.We shall discuss the extension of the Green’s function method to the direct solutionof problems with inhomogeneous boundary conditions in subsection 21.5.2,but we first highlight how the Green’s function approach to solving ODEs canbe simply extended to PDEs <strong>for</strong> homogeneous boundary conditions.21.5.1 Similarities to Green’s functions <strong>for</strong> ODEsAs in the discussion of ODEs in chapter 15, we may consider the Green’sfunction <strong>for</strong> a system described by a PDE as the response of the system to a ‘unitimpulse’ or ‘point source’. Thus if we seek a solution to (21.76) that satisfies somehomogeneous boundary conditions on u(r) then the Green’s function G(r, r 0 )<strong>for</strong>the problem is a solution ofLG(r, r 0 )=δ(r − r 0 ), (21.77)where r 0 lies in V . The Green’s function G(r, r 0 ) must also satisfy the imposed(homogeneous) boundary conditions.It is understood that in (21.77) the L operator expresses differentiation withrespect to r as opposed to r 0 . Also, δ(r − r 0 ) is the Dirac delta function (seechapter 13) of dimension appropriate to the problem; it may be thought of asrepresenting a unit-strength point source at r = r 0 .Following an analogous argument to that given in subsection 15.2.5 <strong>for</strong> ODEs,752


21.5 INHOMOGENEOUS PROBLEMS – GREEN’S FUNCTIONSif the boundary conditions on u(r) are homogeneous then a solution to (21.76)that satisfies the imposed boundary conditions is given by∫u(r) = G(r, r 0 )ρ(r 0 ) dV (r 0 ), (21.78)where the integral on r 0 is over some appropriate ‘volume’. In two or moredimensions, however, the task of finding directly a solution to (21.77) that satisfiesthe imposed boundary conditions on S can be a difficult one, <strong>and</strong> we return tothis in the next subsection.An alternative approach is to follow a similar argument to that presented inchapter 17 <strong>for</strong> ODEs <strong>and</strong> so to construct the Green’s function <strong>for</strong> (21.76) as asuperposition of eigenfunctions of the operator L, provided L is Hermitian. Byanalogy with an ordinary differential operator, a partial differential operator isHermitian if it satisfies∫[∫] ∗v ∗ (r)Lw(r) dV = w ∗ (r)Lv(r) dV ,VVwhere the asterisk denotes complex conjugation <strong>and</strong> v <strong>and</strong> w are arbitrary functionsobeying the imposed (homogeneous) boundary condition on the solution ofLu(r) =0.The eigenfunctions u n (r), n =0, 1, 2,...,ofL satisfyLu n (r) =λ n u n (r),where λ n are the corresponding eigenvalues, which are all real <strong>for</strong> an Hermitianoperator L. Furthermore, each eigenfunction must obey any imposed (homogeneous)boundary conditions. Using an argument analogous to that given inchapter 17, the Green’s function <strong>for</strong> the problem is given by∞∑ u n (r)u ∗G(r, r 0 )=n(r 0 ). (21.79)λ nn=0From (21.79) we see immediately that the Green’s function (irrespective of howit is found) enjoys the propertyG(r, r 0 )=G ∗ (r 0 , r).Thus, if the Green’s function is real then it is symmetric in its two arguments.Once the Green’s function has been obtained, the solution to (21.76) is againgiven by (21.78). For PDEs this approach can become very cumbersome, however,<strong>and</strong> so we shall not pursue it further here.21.5.2 General boundary-value problemsAs mentioned above, often inhomogeneous boundary conditions can be dealtwith by making an appropriate change of variables, such that the boundary753


PDES: SEPARATION OF VARIABLES AND OTHER METHODSconditions in the new variables are homogeneous although the equation itself isgenerally inhomogeneous. In this section, however, we extend the use of Green’sfunctions to problems with inhomogeneous boundary conditions (<strong>and</strong> equations).This provides a more consistent <strong>and</strong> intuitive approach to the solution of suchboundary-value problems.For definiteness we shall consider Poisson’s equation∇ 2 u(r) =ρ(r), (21.80)but the material of this section may be extended to other linear PDEs of the <strong>for</strong>m(21.76). Clearly, Poisson’s equation reduces to Laplace’s equation <strong>for</strong> ρ(r) =0<strong>and</strong>so our discussion is equally applicable to this case.We wish to solve (21.80) in some region V bounded by a surface S, which mayconsist of several disconnected parts. As stated above, we shall allow the possibilitythat the boundary conditions on the solution u(r) may be inhomogeneous on S,although as we shall see this method reduces to those discussed above in thespecial case that the boundary conditions are in fact homogeneous.The two common types of inhomogeneous boundary condition <strong>for</strong> Poisson’sequation are (as discussed in subsection 20.6.2):(i) Dirichlet conditions, in which u(r) is specified on S, <strong>and</strong>(ii) Neumann conditions, in which ∂u/∂n is specified on S.In general, specifying both Dirichlet <strong>and</strong> Neumann conditions on S overdeterminesthe problem <strong>and</strong> leads to there being no solution.The specification of the surface S requires some further comment, since Smay have several disconnected parts. If we wish to solve Poisson’s equationinside some closed surface S then the situation is straight<strong>for</strong>ward <strong>and</strong> is shownin figure 21.11(a). If, however, we wish to solve Poisson’s equation in the gapbetween two closed surfaces (<strong>for</strong> example in the gap between two concentricconducting cylinders) then the volume V is bounded by a surface S that hastwo disconnected parts S 1 <strong>and</strong> S 2 , as shown in figure 21.11(b); the direction ofthe normal to the surface is always taken as pointing out of the volume V .Asimilar situation arises when we wish to solve Poisson’s equation outside someclosed surface S 1 . In this case the volume V is infinite but is treated <strong>for</strong>mallyby taking the surface S 2 as a large sphere of radius R <strong>and</strong> letting R tend toinfinity.In order to solve (21.80) subject to either Dirichlet or Neumann boundaryconditions on S, we will remind ourselves of Green’s second theorem, equation(11.20), which states that, <strong>for</strong> two scalar functions φ(r) <strong>and</strong>ψ(r) defined in somevolume V bounded by a surface S,∫∫(φ∇ 2 ψ − ψ∇ 2 φ) dV = (φ∇ψ − ψ∇φ) · ˆn dS, (21.81)V754S


21.5 INHOMOGENEOUS PROBLEMS – GREEN’S FUNCTIONSVVSˆnS 1S 2ˆnˆn(a)Figure 21.11regions V .Surfaces used <strong>for</strong> solving Poisson’s equation in different(b)where on the RHS it is common to write, <strong>for</strong> example, ∇ψ · ˆn dS as (∂ψ/∂n) dS.The expression ∂ψ/∂n st<strong>and</strong>s <strong>for</strong> ∇ψ · ˆn, the rate of change of ψ in the directionof the unit outward normal ˆn to the surface S.The Green’s function <strong>for</strong> Poisson’s equation (21.80) must satisfy∇ 2 G(r, r 0 )=δ(r − r 0 ), (21.82)where r 0 lies in V . (As mentioned above, we may think of G(r, r 0 )asthesolutionto Poisson’s equation <strong>for</strong> a unit-strength point source located at r = r 0 .) Let us<strong>for</strong> the moment impose no boundary conditions on G(r, r 0 ).If we now let φ = u(r) <strong>and</strong>ψ = G(r, r 0 ) in Green’s theorem (21.81) then weobtain∫[u(r)∇ 2 G(r, r 0 ) − G(r, r 0 ) ∇ 2 u(r) ] dV (r)V∫ [= u(r) ∂G(r, r 0)− G(r, r 0 ) ∂u(r) ]dS(r),S ∂n∂nwhere we have made explicit that the volume <strong>and</strong> surface integrals are withrespect to r. Using (21.80) <strong>and</strong> (21.82) the LHS can be simplified to give∫[u(r)δ(r − r 0 ) − G(r, r 0 )ρ(r)] dV (r)V∫ [= u(r) ∂G(r, r 0)− G(r, r 0 ) ∂u(r) ]dS(r). (21.83)S ∂n∂nSince r 0 lies within the volume V ,∫u(r)δ(r − r 0 ) dV (r) =u(r 0 ),V<strong>and</strong> thus on rearranging (21.83) the solution to Poisson’s equation (21.80) can be755


PDES: SEPARATION OF VARIABLES AND OTHER METHODSwritten as∫u(r 0 )=V∫ [G(r, r 0 )ρ(r) dV (r)+ u(r) ∂G(r, r 0)− G(r, r 0 ) ∂u(r) ]dS(r).S ∂n∂n(21.84)Clearly, we can interchange the roles of r <strong>and</strong> r 0 in (21.84) if we wish. (Rememberalso that, <strong>for</strong> a real Green’s function, G(r, r 0 )=G(r 0 , r).)Equation (21.84) is central to the extension of the Green’s function methodto problems with inhomogeneous boundary conditions, <strong>and</strong> we next discuss itsapplication to both Dirichlet <strong>and</strong> Neumann boundary-value problems. But, be<strong>for</strong>edoing so, we also note that if the boundary condition on S is in fact homogeneous,so that u(r) =0or∂u(r)/∂n =0onS, then dem<strong>and</strong>ing that the Green’s functionG(r, r 0 ) also obeys the same boundary condition causes the surface integral in(21.84) to vanish, <strong>and</strong> we are left with the familiar <strong>for</strong>m of solution given in(21.78). The extension of (21.84) to a PDE other than Poisson’s equation isdiscussed in exercise 21.28.21.5.3 Dirichlet problemsIn a Dirichlet problem we require the solution u(r) of Poisson’s equation (21.80)to take specific values on some surface S that bounds V ,i.e.werequirethatu(r) =f(r) onS where f is a given function.If we seek a Green’s function G(r, r 0 ) <strong>for</strong> this problem it must clearly satisfy(21.82), but we are free to choose the boundary conditions satisfied by G(r, r 0 )insuch a way as to make the solution (21.84) as simple as possible. From (21.84),we see that by choosingG(r, r 0 )=0 <strong>for</strong>r on S (21.85)the second term in the surface integral vanishes. Since u(r) =f(r) onS, (21.84)then becomes∫∫u(r 0 )= G(r, r 0 )ρ(r) dV (r)+ f(r) ∂G(r, r 0)dS(r). (21.86)VS ∂nThus we wish to find the Dirichlet Green’s function that(i) satisfies (21.82) <strong>and</strong> hence is singular at r = r 0 ,<strong>and</strong>(ii) obeys the boundary condition G(r, r 0 )=0<strong>for</strong>r on S.In general, it is difficult to obtain this function directly, <strong>and</strong> so it is useful toseparate these two requirements. We there<strong>for</strong>e look <strong>for</strong> a solution of the <strong>for</strong>mG(r, r 0 )=F(r, r 0 )+H(r, r 0 ),where F(r, r 0 ) satisfies (21.82) <strong>and</strong> has the required singular character at r = r 0 butdoes not necessarily obey the boundary condition on S, whilst H(r, r 0 ) satisfies756


21.5 INHOMOGENEOUS PROBLEMS – GREEN’S FUNCTIONSthe corresponding homogeneous equation (i.e. Laplace’s equation) inside V butis adjusted in such a way that the sum G(r, r 0 ) equals zero on S. The Green’sfunction G(r, r 0 ) is still a solution of (21.82) since∇ 2 G(r, r 0 )=∇ 2 F(r, r 0 )+∇ 2 H(r, r 0 )=∇ 2 F(r, r 0 )+0=δ(r − r 0 ).The function F(r, r 0 ) is called the fundamental solution <strong>and</strong> will clearly takedifferent <strong>for</strong>ms depending on the dimensionality of the problem. Let us firstconsider the fundamental solution to (21.82) in three dimensions.◮Find the fundamental solution to Poisson’s equation in three dimensions that tends to zeroas |r| →∞.We wish to solve∇ 2 F(r, r 0 )=δ(r − r 0 ) (21.87)in three dimensions, subject to the boundary condition F(r, r 0 ) → 0as|r| →∞.Sincetheproblem is spherically symmetric about r 0 , let us consider a large sphere S of radius Rcentred on r 0 , <strong>and</strong> integrate (21.87) over the enclosed volume V . We then obtain∫∫∇ 2 F(r, r 0 ) dV = δ(r − r 0 ) dV =1, (21.88)Vsince V encloses the point r 0 . However, using the divergence theorem,∫∫∇ 2 F(r, r 0 ) dV = ∇F(r, r 0 ) · ˆn dS, (21.89)Vwhere ˆn is the unit normal to the large sphere S at any point.Since the problem is spherically symmetric about r 0 , we expect thatF(r, r 0 )=F(|r − r 0 |)=F(r),i.e. that F has the same value everywhere on S. Thus, evaluating the surface integral in(21.89) <strong>and</strong> equating it to unity from (21.88), we have §4πr 2 dFdr ∣ =1.r=RIntegrating this expression we obtainF(r) =− 14πr +constant,but, since we require F(r, r 0 ) → 0as|r| →∞, the constant must be zero. The fundamentalsolution in three dimensions is consequently given by1F(r, r 0 )=−4π|r − r 0 | . (21.90)This is clearly also the full Green’s function <strong>for</strong> Poisson’s equation subject to the boundarycondition u(r) → 0as|r| →∞. ◭VUsing (21.90) we can write down the solution of Poisson’s equation to find,S§ A vertical bar to the right of an expression is a common alternative to enclosing the expression insquare brackets; as usual, the subscript shows the value of the variable at which the expression isto be evaluated.757


PDES: SEPARATION OF VARIABLES AND OTHER METHODS<strong>for</strong> example, the electrostatic potential u(r) due to some distribution of electriccharge ρ(r). The electrostatic potential satisfies∇ 2 u(r) =− ρ ɛ 0,where u(r) → 0as|r| →∞. Since the boundary condition on the surface atinfinity is homogeneous the surface integral in (21.86) vanishes, <strong>and</strong> using (21.90)we recover the familiar solution∫u(r 0 )=ρ(r)dV (r), (21.91)4πɛ 0 |r − r 0 |where the volume integral is over all space.We can develop an analogous theory in two dimensions. As be<strong>for</strong>e the fundamentalsolution satisfies∇ 2 F(r, r 0 )=δ(r − r 0 ), (21.92)where δ(r−r 0 ) is now the two-dimensional delta function. Following an analogousmethod to that used in the previous example, we find the fundamental solutionin two dimensions to be given byF(r, r 0 )= 12π ln |r − r 0| + constant. (21.93)From the <strong>for</strong>m of the solution we see that in two dimensions we cannot applythe condition F(r, r 0 ) → 0as|r| →∞, <strong>and</strong> in this case the constant does notnecessarily vanish.We now return to the task of constructing the full Dirichlet Green’s function. Todo so we wish to add to the fundamental solution a solution of the homogeneousequation (in this case Laplace’s equation) such that G(r, r 0 )=0onS, asrequiredby (21.86) <strong>and</strong> its attendant conditions. The appropriate Green’s function isconstructed by adding to the fundamental solution ‘copies’ of itself that represent‘image’ sources at different locations outside V . Hence this approach is called themethod of images.In summary, if we wish to solve Poisson’s equation in some region V subject toDirichlet boundary conditions on its surface S then the procedure <strong>and</strong> argumentare as follows.(i) To the single source δ(r − r 0 ) inside V add image sources outside VN∑q n δ(r − r n ) with r n outside V,n=1where the positions r n <strong>and</strong> the strengths q n of the image sources are to bedetermined as described in step (iii) below.758


21.5 INHOMOGENEOUS PROBLEMS – GREEN’S FUNCTIONS(ii) Since all the image sources lie outside V , the fundamental solution correspondingto each source satisfies Laplace’s equation inside V . Thus wemay add the fundamental solutions F(r, r n ) corresponding to each imagesource to that corresponding to the single source inside V , obtaining theGreen’s functionN∑G(r, r 0 )=F(r, r 0 )+ q n F(r, r n ).(iii) Now adjust the positions r n <strong>and</strong> strengths q n of the image sources sothat the required boundary conditions are satisfied on S. For a DirichletGreen’s function we require G(r, r 0 )=0<strong>for</strong>r on S.(iv) The solution to Poisson’s equation subject to the Dirichlet boundarycondition u(r) =f(r) onS is then given by (21.86).In general it is very difficult to find the correct positions <strong>and</strong> strengths <strong>for</strong> theimages, i.e. to make them such that the boundary conditions on S are satisfied.Nevertheless, it is possible to do so <strong>for</strong> certain problems that have simple geometry.In particular, <strong>for</strong> problems in which the boundary S consists of straight lines (intwo dimensions) or planes (in three dimensions), positions of the image pointscan be deduced simply by imagining the boundary lines or planes to be mirrorsin which the single source in V (at r 0 ) is reflected.◮Solve Laplace’s equation ∇ 2 u =0in three dimensions in the half-space z>0, given thatu(r) =f(r) on the plane z =0.The surface S bounding V consists of the xy-plane <strong>and</strong> the surface at infinity. There<strong>for</strong>e,the Dirichlet Green’s function <strong>for</strong> this problem must satisfy G(r, r 0 )=0onz =0<strong>and</strong>G(r, r 0 ) → 0as|r| →∞. Thus it is clear in this case that we require one image source at aposition r 1 that is the reflection of r 0 in the plane z = 0, as shown in figure 21.12 (so thatr 1 lies in z


PDES: SEPARATION OF VARIABLES AND OTHER METHODSzV+r 0yx−r 1Figure 21.12 The arrangement of images <strong>for</strong> solving Laplace’s equation inthe half-space z>0.We may evaluate this normal derivative by writing the Green’s function (21.94) explicitlyin terms of x, y <strong>and</strong> z (<strong>and</strong> x 0 , y 0 <strong>and</strong> z 0 ) <strong>and</strong> calculating the partial derivative with respectto z directly. It is usually quicker, however, to use the fact that §∇|r − r 0 | = r − r 0|r − r 0 | ; (21.96)thus∇G(r, r 0 )= r − r 04π|r − r 0 | − r − r 13 4π|r − r 1 | . 3Since r 0 =(x 0 ,y 0 ,z 0 )<strong>and</strong>r 1 =(x 0 ,y 0 , −z 0 ) the normal derivative is given by− ∂G(r, r 0)= −k · ∇G(r, r 0 )∂z= − z − z 04π|r − r 0 | + z + z 03 4π|r − r 1 | . 3There<strong>for</strong>e on the surface z = 0, <strong>and</strong> writing out the dependence on x, y <strong>and</strong> z explicitly,we have− ∂G(r, r 0)2z 0∂z ∣ =z=04π[(x − x 0 ) 2 +(y − y 0 ) 2 + z .0 2]3/2 Inserting this expression into (21.95) we obtain the solutionu(x 0 ,y 0 ,z 0 )= z ∫ ∞ ∫ ∞0f(x, y)dx dy. ◭2π −∞ −∞ [(x − x 0 ) 2 +(y − y 0 ) 2 + z0 2 ]3/2An analogous procedure may be applied in two-dimensional problems. For§ Since |r − r 0 | 2 =(r − r 0 ) · (r − r 0 ) we have ∇|r − r 0 | 2 =2(r − r 0 ), from which we obtain∇(|r − r 0 | 2 ) 1/2 = 1 2(r − r 0 )2 (|r − r 0 | 2 ) 1/2 = r − r 0|r − r 0 | .Note that this result holds in two <strong>and</strong> three dimensions.760


21.5 INHOMOGENEOUS PROBLEMS – GREEN’S FUNCTIONSexample, in solving Poisson’s equation in two dimensions in the half-space x>0we again require just one image charge, of strength q 1 = −1, at a position r 1 thatis the reflection of r 0 in the line x = 0. Since we require G(r, r 0 )=0whenr lieson x = 0, the constant in (21.93) must equal zero, <strong>and</strong> so the Dirichlet Green’sfunction isG(r, r 0 )= 1 (ln |r − r0 |−ln |r − r 1 | ) .2πClearly G(r, r 0 ) tends to zero as |r| →∞. If, however, we wish to solve the twodimensionalPoisson equation in the quarter space x>0, y>0, then more imagepoints are required.◮A line charge in the z-direction of charge density λ is placed at some position r 0 in thequarter-space x>0, y>0. Calculate the <strong>for</strong>ce per unit length on the line charge due tothe presence of thin earthed plates along x =0<strong>and</strong> y =0.Here we wish to solve Poisson’s equation,∇ 2 u = − λ ɛ 0δ(r − r 0 ),in the quarter space x>0, y>0. It is clear that we require three image line chargeswith positions <strong>and</strong> strengths as shown in figure 21.13 (all of which lie outside the regionin which we seek a solution). The boundary condition that the electrostatic potential u iszero on x =0<strong>and</strong>y = 0 (shown as the ‘curve’ C in figure 21.13) is then automaticallysatisfied, <strong>and</strong> so this system of image charges is directly equivalent to the original situationof a single line charge in the presence of the earthed plates along x =0<strong>and</strong>y = 0. Thusthe electrostatic potential is simply equal to the Dirichlet Green’s functionu(r) =G(r, r 0 )=−λ (ln |r − r0 |−ln |r − r 1 | +ln|r − r 2 |−ln |r − r 3 | ) ,2πɛ 0which equals zero on C <strong>and</strong> on the ‘surface’ at infinity.The <strong>for</strong>ce on the line charge at r 0 , there<strong>for</strong>e, is simply that due to the three line chargesat r 1 , r 2 <strong>and</strong> r 3 . The elecrostatic potential due to a line charge at r i , i = 1, 2 or 3, is givenby the fundamental solutionu i (r) =∓λ ln |r − r i | + c,2πɛ 0the upper or lower sign being taken according to whether the line charge is positive ornegative, respectively. There<strong>for</strong>e the <strong>for</strong>ce per unit length on the line charge at r 0 , due tothe one at r i ,isgivenby−λ∇u i (r)∣ = ± λ2 r 0 − r ir=r02πɛ 0 |r 0 − r i | . 2Adding the contributions from the three image charges shown in figure 21.13, the total<strong>for</strong>ce experienced by the line charge at r 0 is given by(F =λ2− r 0 − r 12πɛ 0 |r 0 − r 1 | + r 0 − r 22 |r 0 − r 2 | − r )0 − r 3,2 |r 0 − r 3 | 2where, from the figure, r 0 − r 1 =2y 0 j, r 0 − r 2 =2x 0 i +2y 0 j <strong>and</strong> r 0 − r 3 =2x 0 i. Thus, in761


PDES: SEPARATION OF VARIABLES AND OTHER METHODSy−λ+λr 0r 3C xx 0y 0V+λr 2−λr 1Figure 21.13 The arrangement of images <strong>for</strong> finding the <strong>for</strong>ce on a linecharge situated in the (two-dimensional) quarter-space x>0, y>0, when theplanes x =0<strong>and</strong>y = 0 are earthed.terms of x 0 <strong>and</strong> y 0 , the total <strong>for</strong>ce on the line charge due to the charge induced on theplates is given by(F =λ2− 1 j + 2x 0 i +2y 0 j− 1 )i2πɛ 0 2y 0 4x 2 0 +4y2 02x 0( )λ 2 y2= −04πɛ 0 (x 2 0 + y2 0 ) i + x2 0j . ◭x 0 y 0Further generalisations are possible. For instance, solving Poisson’s equation inthe two-dimensional strip −∞


21.5 INHOMOGENEOUS PROBLEMS – GREEN’S FUNCTIONSVza+1 r 0A−a|r 0 | r 1xyB−aFigure 21.14 The arrangement of images <strong>for</strong> solving Poisson’s equationoutside a sphere of radius a centred at the origin. For a charge +1 at r 0 ,theimage point r 1 is given by (a/|r 0 |) 2 r 0 <strong>and</strong> the strength of the image charge is−a/|r 0 |.By symmetry we expect the image point r 1 to lie on the same radial line as the originalsource, r 0 , as shown in figure 21.14, <strong>and</strong> so r 1 = kr 0 where k


PDES: SEPARATION OF VARIABLES AND OTHER METHODSThere<strong>for</strong>e, in order that G =0at|r| = a, the strength of the image charge must be−a/|r 0 |. Consequently, the Dirichlet Green’s function <strong>for</strong> the exterior of the sphere is1G(r, r 0 )=−4π|r − r 0 | + a/|r 0 |4π |r − (a 2 /|r 0 | 2 )r 0 | .For a less <strong>for</strong>mal treatment of the same problem see exercise 21.22. ◭If we seek solutions to Poisson’s equation in the interior of a sphere then theabove analysis still holds, but r <strong>and</strong> r 0 are now inside the sphere <strong>and</strong> the imager 1 lies outside it.For two-dimensional Dirichlet problems outside the circle |r| = a, we are ledby arguments similar to those employed previously to use the same image pointas in the three-dimensional case, namelyr 1 =a2|r 0 | 2 r 0. (21.100)As illustrated below, however, it is usually necessary to take the image strengthas −1 in two-dimensional problems.◮Solve Laplace’s equation in the two-dimensional region |r| ≤a, subject to the boundarycondition u = f(φ) on |r| = a.In this case we wish to find the Dirichlet Green’s function in the interior of a disc ofradius a, so the image charge must lie outside the disc. Taking the strength of the imageto be −1, we haveG(r, r 0 )= 12π ln |r − r 0|− 12π ln |r − r 1| + c,where r 1 =(a 2 /|r 0 | 2 )r 0 lies outside the disc, <strong>and</strong> c is a constant that includes the strengthof the image charge <strong>and</strong> does not necessarily equal zero.Since we require G(r, r 0 )=0when|r| = a, the value of the constant c is determined,<strong>and</strong> the Dirichlet Green’s function <strong>for</strong> this problem is given byG(r, r 0 )= 1 (ln |r − r 0 |−ln2π∣ r −a2|r 0 | r 2 0∣ − ln |r )0|. (21.101)aUsing plane polar coordinates, the solution to the boundary-value problem can be writtenas a line integral around the circle ρ = a:∫u(r 0 )= f(r) ∂G(r, r 0)dlC ∂n∫ 2π= f(r) ∂G(r, r 0)0 ∂ρ ∣ adφ. (21.102)ρ=aThe normal derivative of the Green’s function (21.101) is given by∂G(r, r 0 )∂ρ= r|r| · ∇G(r, r 0)= r2π|r| ·( r − r0|r − r 0 | − r − r )1. (21.103)2 |r − r 1 | 2764


21.5 INHOMOGENEOUS PROBLEMS – GREEN’S FUNCTIONSUsing the fact that r 1 =(a 2 /|r 0 | 2 )r 0 <strong>and</strong> the geometrical result (21.99), we find that∂G(r, r 0 )∂ρ ∣ = a2 −|r 0 | 2ρ=a2πa|r − r 0 | . 2In plane polar coordinates, r = ρ cos φ i + ρ sin φ j <strong>and</strong> r 0 = ρ 0 cos φ 0 i + ρ 0 sin φ 0 j,<strong>and</strong> so( )∂G(r, r 0 )1a 2 − ρ 2 0∂ρ ∣ =ρ=a2πa a 2 + ρ 2 0 − 2aρ 0 cos(φ − φ 0 ) .On substituting into (21.102), we obtain∫ 2πu(ρ 0 ,φ 0 )= 12π 0which is the solution to the problem. ◭(a 2 − ρ 2 0 )f(φ) dφa 2 + ρ 2 0 − 2aρ 0 cos(φ − φ 0 ) , (21.104)21.5.4 Neumann problemsIn a Neumann problem we require the normal derivative of the solution ofPoisson’s equation to take on specific values on some surface S that bounds V ,i.e. we require ∂u(r)/∂n = f(r) onS, wheref is a given function. As we shall see,much of our discussion of Dirichlet problems can be immediately taken over intothe solution of Neumann problems.As we proved in section 20.7 of the previous chapter, specifying Neumannboundary conditions determines the relevant solution of Poisson’s equation towithin an (unimportant) additive constant. Unlike Dirichlet conditions, Neumannconditions impose a self-consistency requirement. In order <strong>for</strong> a solution u to exist,it is necessary that the following consistency condition holds:∫ ∫∫∫fdS = ∇u · ˆn dS = ∇ 2 udV = ρdV, (21.105)SSwhere we have used the divergence theorem to convert the surface integral intoa volume integral. As a physical example, the integral of the normal componentof an electric field over a surface bounding a given volume cannot be chosenarbitrarily when the charge inside the volume has already been specified (Gauss’stheorem).Let us again consider (21.84), which is central to our discussion of Green’sfunctions in inhomogeneous problems. It reads∫∫ [u(r 0 )= G(r, r 0 )ρ(r) dV (r)+ u(r) ∂G(r, r 0)− G(r, r 0 ) ∂u(r) ]dS(r).VS ∂n∂nAs always, the Green’s function must obeyV∇ 2 G(r, r 0 )=δ(r − r 0 ),where r 0 lies in V . In the solution of Dirichlet problems in the previous subsection,we chose the Green’s function to obey the boundary condition G(r, r 0 )=0onS765V


PDES: SEPARATION OF VARIABLES AND OTHER METHODS<strong>and</strong>, in a similar way, we might wish to choose ∂G(r, r 0 )/∂n = 0 in the solution ofNeumann problems. However, in general this is not permitted since the Green’sfunction must obey the consistency condition∫∫∂G(r, r 0 )dS = ∇G(r, r 0 ) · ˆn dS = ∇ 2 G(r, r 0 ) dV =1.S ∂nSVThe simplest permitted boundary condition is there<strong>for</strong>e∂G(r, r 0 )= 1 <strong>for</strong> r on S,∂n Awhere A is the area of the surface S; this defines a Neumann Green’s function.If we require ∂u(r)/∂n = f(r) onS, the solution to Poisson’s equation is given by∫u(r 0 )= G(r, r 0 )ρ(r) dV (r)+ 1 ∫∫u(r) dS(r) − G(r, r 0 )f(r) dS(r)V A SS∫∫= G(r, r 0 )ρ(r) dV (r)+〈u(r)〉 S − G(r, r 0 )f(r) dS(r), (21.106)Vwhere 〈u(r)〉 S is the average of u over the surface S <strong>and</strong> is a freely specifiable constant.For Neumann problems in which the volume V is bounded by a surface S atinfinity, we do not need the 〈u(r)〉 S term. For example, if we wish to solve a Neumannproblem outside the unit sphere centred at the origin then r>ais the regionV throughout which we require the solution; this region may be considered as beingbounded by two disconnected surfaces, the surface of the sphere <strong>and</strong> a surfaceat infinity. By requiring that u(r) → 0as|r| →∞,theterm〈u(r)〉 S becomes zero.As mentioned above, much of our discussion of Dirichlet problems can betaken over into the solution of Neumann problems. In particular, we may use themethod of images to find the appropriate Neumann Green’s function.◮Solve Laplace’s equation in the two-dimensional region |r| ≤a subject to the boundarycondition ∂u/∂n = f(φ) on |r| = a, with ∫ 2πf(φ) dφ =0as required by the consistency0condition (21.105).Let us assume, as in Dirichlet problems with this geometry, that a single image charge isplaced outside the circle atr 1 =a2|r 0 | r 0,2where r 0 is the position of the source inside the circle (see equation (21.100)). Then, from(21.99), we have the useful geometrical result|r − r 1 | = a|r 0 | |r − r 0| <strong>for</strong> |r| = a. (21.107)Leaving the strength q of the image as a parameter, the Green’s function has the <strong>for</strong>mG(r, r 0 )= 1 (ln |r − r0 | + q ln |r − r 1 | + c ) . (21.108)2π766S∫


21.6 EXERCISESUsing plane polar coordinates, the radial (i.e. normal) derivative of this function is givenby∂G(r, r 0 )∂ρ= r|r| · ∇G(r, r 0)= r2π|r| ·[ r − r0|r − r 0 | + q(r − r ]1).2 |r − r 1 | 2Using (21.107), on the perimeter of the circle ρ = a the radial derivative takes the <strong>for</strong>m∂G(r, r 0 )∂ρ ∣ = 1 [ ]|r| 2 − r · r 0+ q|r|2 − q(a 2 /|r 0 | 2 )r · r 0ρ=a2π|r| |r − r 0 | 2 (a 2 /|r 0 | 2 )|r − r 0 | 2= 1 1 [ ]|r| 2 + q|r2πa |r − r 0 | 2 0 | 2 − (1 + q)r · r 0 ,where we have set |r| 2 = a 2 in the second term on the RHS, but not in the first. If we takeq = 1, the radial derivative simplifies to∂G(r, r 0 )∂ρ ∣ = 1ρ=a2πa ,or 1/L, whereL is the circumference, <strong>and</strong> so (21.108) with q = 1 is the required NeumannGreen’s function.Since ρ(r) = 0, the solution to our boundary-value problem is now given by (21.106) as∫u(r 0 )=〈u(r)〉 C − G(r, r 0 )f(r) dl(r),where the integral is around the circumference of the circle C. In plane polar coordinatesr = ρ cos φ i + ρ sin φ j <strong>and</strong> r 0 = ρ 0 cos φ 0 i + ρ 0 sin φ 0 j, <strong>and</strong> again using (21.107) we findthat on C the Green’s function is given byG(r, r 0 )| ρ=a= 12π= 12π= 12πC[( ) ]aln |r − r 0 | +ln|r 0 | |r − r 0| + c(ln |r − r 0 | 2 +ln a )|r 0 | + c {ln [ a 2 + ρ 2 0 − 2aρ 0 cos(φ − φ 0 ) ] +ln a }+ c . (21.109)ρ 0Since dl = adφ on C, the solution to the problem is given byu(ρ 0 ,φ 0 )=〈u〉 C − a ∫ 2πf(φ)ln[a 2 + ρ 2 0 − 2aρ 0 cos(φ − φ 0 )] dφ.2π 0The contributions of the final two terms terms in the Green’s function (21.109) vanishbecause ∫ 2πf(φ) dφ = 0. The average value of u around the circumference, 〈u〉0 C , is a freelyspecifiable constant as we would expect <strong>for</strong> a Neumann problem. This result should becompared with the result (21.104) <strong>for</strong> the corresponding Dirichlet problem, but it shouldbe remembered that in the one case f(φ) is a potential, <strong>and</strong> in the other the gradient of apotential. ◭21.6 Exercises21.1 Solve the following first-order partial differential equations by separating thevariables:(a) ∂u∂x − x ∂u∂y =0;∂u ∂u(b)x − 2y∂x ∂y =0.767


PDES: SEPARATION OF VARIABLES AND OTHER METHODS21.2 A cube, made of material whose conductivity is k, has as its six faces the planesx = ±a, y = ±a <strong>and</strong> z = ±a, <strong>and</strong> contains no internal heat sources. Verify thatthe temperature distributionu(x, y, z, t) =A cos πx ( )πzsina a exp − 2κπ2 ta 2obeys the appropriate diffusion equation. Across which faces is there heat flow?What is the direction <strong>and</strong> rate of heat flow at the point (3a/4,a/4,a)attimet = a 2 /(κπ 2 )?21.3 The wave equation describing the transverse vibrations of a stretched membraneunder tension T <strong>and</strong> having a uni<strong>for</strong>m surface density ρ is( )∂ 2 uT∂x + ∂2 u= ρ ∂2 u2 ∂y 2 ∂t . 2Find a separable solution appropriate to a membrane stretched on a frame oflength a <strong>and</strong> width b, showing that the natural angular frequencies of such amembrane are given by( )ω 2 = π2 T n2ρ a + m2,2 b 2where n <strong>and</strong> m are any positive integers.21.4 Schrödinger’s equation <strong>for</strong> a non-relativistic particle in a constant potential regioncan be taken as( )− 2 ∂ 2 u2m ∂x + ∂2 u2 ∂y + ∂2 u= i ∂u2 ∂z 2 ∂t .(a) Find a solution, separable in the four independent variables, that can bewritten in the <strong>for</strong>m of a plane wave,ψ(x, y, z, t) =A exp[i(k · r − ωt)].Using the relationships associated with de Broglie (p = k) <strong>and</strong> Einstein(E = ω), show that the separation constants must be such thatp 2 x + p 2 y + p 2 z =2mE.(b) Obtain a different separable solution describing a particle confined to a boxof side a (ψ must vanish at the walls of the box). Show that the energy ofthe particle can only take the quantised valuesE = 2 π 22ma 2 (n2 x + n 2 y + n 2 z),where n x , n y <strong>and</strong> n z are integers.21.5 Denoting the three terms of ∇ 2 in spherical polars by ∇ 2 r, ∇ 2 θ , ∇2 φin an obviousway, evaluate ∇ 2 r u, etc. <strong>for</strong> the two functions given below <strong>and</strong> verify that, in eachcase, although the individual terms are not necessarily zero their sum ∇ 2 u is zero.Identify the corresponding values of l <strong>and</strong> m.((a) u(r, θ, φ) = Ar 2 + B ) 3cos 2 θ − 1.r 3 2(b) u(r, θ, φ) =(Ar + B )sin θ exp iφ.r 221.6 Prove that the expression given in equation (21.47) <strong>for</strong> the associated Legendrefunction Pl m (µ) satisfies the appropriate equation, (21.45), as follows.768


21.6 EXERCISES(a) Evaluate dPl m(µ)/dµ <strong>and</strong> d2 Pl m(µ)/dµ2 , using the <strong>for</strong>ms given in (21.47), <strong>and</strong>substitute them into (21.45).(b) Differentiate Legendre’s equation m times using Leibnitz’ theorem.(c) Show that the equations obtained in (a) <strong>and</strong> (b) are multiples of each other,<strong>and</strong> hence that the validity of (b) implies that of (a).21.7 Continue the analysis of exercise 10.20, concerned with the flow of a very viscousfluid past a sphere, to find the full expression <strong>for</strong> the stream function ψ(r, θ). Atthe surface of the sphere r = a, the velocity field u = 0, whilst far from the sphereψ ≃ (Ur 2 sin 2 θ)/2.Show that f(r) can be expressed as a superposition of powers of r, <strong>and</strong>determine which powers give acceptable solutions. Hence show thatψ(r, θ) = U )(2r 2 − 3ar + a3sin 2 θ.4r21.8 The motion of a very viscous fluid in the two-dimensional (wedge) region −α


PDES: SEPARATION OF VARIABLES AND OTHER METHODSx = 0, <strong>and</strong> free at the other, x = L, show that the angular frequency of vibrationω satisfies( ) ( )ω 1/2 Lω 1/2 Lcosh = − sec .aa[ At a clamped end both u <strong>and</strong> ∂u/∂x vanish, whilst at a free end, where there isno bending moment, ∂ 2 u/∂x 2 <strong>and</strong> ∂ 3 u/∂x 3 are both zero. ]21.12 A membrane is stretched between two concentric rings of radii a <strong>and</strong> b (b >a).If the smaller ring is transversely distorted from the planar configuration by anamount c|φ|, −π ≤ φ ≤ π, show that the membrane then has a shape given byu(ρ, φ) = cπ ln(b/ρ)2 ln(b/a) − 4c ∑( )a m b2mπ m 2 (b 2m − a 2m ) ρ − m ρm cos mφ.m odd21.13 A string of length L, fixed at its two ends, is plucked at its mid-point by anamount A <strong>and</strong> then released. Prove that the subsequent displacement is given by∞∑[ ] [ ]8A (−1) n (2n +1)πx (2n +1)πctu(x, t) =π 2 (2n +1) sin cos,2 LLn=0where, in the usual notation, c 2 = T/ρ.Find the total kinetic energy of the string when it passes through its unpluckedposition, by calculating it in each mode (each n) <strong>and</strong> summing, using the result∞∑ 1(2n +1) = π22 8 .0Confirm that the total energy is equal to the work done in plucking the stringinitially.21.14 Prove that the potential <strong>for</strong> ρ0<strong>and</strong>cosφ


21.6 EXERCISES21.18 A sphere of radius a <strong>and</strong> thermal conductivity k 1 is surrounded by an infinitemedium of conductivity k 2 in which far away the temperature tends to T ∞ .A distribution of heat sources q(θ) embedded in the sphere’s surface establishsteady temperature fields T 1 (r, θ) inside the sphere <strong>and</strong> T 2 (r, θ) outside it. It canbe shown, by considering the heat flow through a small volume that includespart of the sphere’s surface, that∂T 1k 1∂r − k ∂T 22 = q(θ) on r = a.∂rGiven thatq(θ) = 1 ∞∑q n P n (cos θ),an=0find complete expressions <strong>for</strong> T 1 (r, θ) <strong>and</strong>T 2 (r, θ). What is the temperature atthe centre of the sphere?21.19 Using result (21.74) from the worked example in the text, find the generalexpression <strong>for</strong> the temperature u(x, t) in the bar, given that the temperaturedistribution at time t =0isu(x, 0) = exp(−x 2 /a 2 ).21.20 Working in spherical polar coordinates r =(r, θ, φ), but <strong>for</strong> a system that hasazimuthal symmetry around the polar axis, consider the following gravitationalproblem.(a) Show that the gravitational potential due to a uni<strong>for</strong>m disc of radius a <strong>and</strong>mass M, centred at the origin, is given <strong>for</strong> raby[GM1 − 1 ( a) 2P2 (cos θ)+ 1 ( a) 4P4 (cos θ) − ···],r 4 r8 rwhere the polar axis is normal to the plane of the disc.(b) Reconcile the presence of a term P 1 (cos θ), which is odd under θ → π − θ,with the symmetry with respect to the plane of the disc of the physicalsystem.(c) Deduce that the gravitational field near an infinite sheet of matter of constantdensity ρ per unit area is 2πGρ.21.21 In the region −∞


PDES: SEPARATION OF VARIABLES AND OTHER METHODS21.22 Point charges q <strong>and</strong> −qa/b (with a0 <strong>for</strong> the solution of∇ 2 Φ = 0 with Φ specified in cylindrical polar coordinates (ρ, φ, z) on the planez =0by{1 <strong>for</strong> ρ ≤ 1,Φ(ρ, φ, z) =1/ρ <strong>for</strong> ρ>1.Determine the variation of Φ(0, 0,z)alongthez-axis.21.24 Electrostatic charge is distributed in a sphere of radius R centred on the origin.Determine the <strong>for</strong>m of the resultant potential φ(r) at distances much greater thanR, as follows.(a) Express in the <strong>for</strong>m of an integral over all space the solution of∇ 2 φ = − ρ(r) .ɛ 0(b) Show that, <strong>for</strong> r ≫ r ′ ,|r − r ′ | = r − r · r′r( ) 1+O .r(c) Use results (a) <strong>and</strong> (b) to show that φ(r) hasthe<strong>for</strong>mφ(r) = M r + d · r ( ) 1+O .r 3 r 3Find expressions <strong>for</strong> M <strong>and</strong> d, <strong>and</strong> identify them physically.21.25 Find, in the <strong>for</strong>m of an infinite series, the Green’s function of the ∇ 2 operator <strong>for</strong>the Dirichlet problem in the region −∞


21.7 HINTS AND ANSWERSin V <strong>and</strong> takes the specified <strong>for</strong>m φ = f on S, the boundary of V .TheGreen’s function, G(r, r ′ ), to be used satisfies∇ 2 G − m 2 G = δ(r − r ′ )<strong>and</strong> vanishes when r is on S.(b) When V is all space, G(r, r ′ ) can be written as G(t) =g(t)/t, wheret = |r − r ′ |<strong>and</strong> g(t) is bounded as t →∞. Find the <strong>for</strong>m of G(t).(c) Find φ(r) in the half-space x>0ifρ(r) =δ(r − r 1 )<strong>and</strong>φ = 0 both on x =0<strong>and</strong> as r →∞.21.28 Consider the PDE Lu(r) =ρ(r), <strong>for</strong> which the differential operator L is given byL = ∇ · [p(r)∇]+q(r),where p(r) <strong>and</strong>q(r) are functions of position. By proving the generalised <strong>for</strong>m ofGreen’s theorem,∫ ∮(φLψ − ψLφ) dV = p(φ∇ψ − ψ∇φ) · ˆn dS,VSshow that the solution of the PDE is given by∫∮ [u(r 0 )= G(r, r 0 )ρ(r) dV (r)+ p(r) u(r) ∂G(r, r 0)− G(r, r 0 ) ∂u(r) ]dS(r),VS∂n∂nwhere G(r, r 0 ) is the Green’s function satisfying LG(r, r 0 )=δ(r − r 0 ).21.7 Hints <strong>and</strong> answers21.1 (a) C exp[λ(x 2 +2y)]; (b) C(x 2 y) λ .21.3 u(x, y, t) = sin(nπx/a) sin(mπy/b)(A sin ωt + B cos ωt).21.5 (a) 6u/r 2 , −6u/r 2 ,0,l =2(or−3), m =0;(b) 2u/r 2 ,(cot 2 θ − 1)u/r 2 ; −u/(r 2 sin 2 θ), l =1(or−2), m = ±1.21.7 Solutions of the <strong>for</strong>m r l give l as −1, 1, 2, 4. Because of the asymptotic <strong>for</strong>m ofψ, anr 4 term cannot be present. The coefficients of the three remaining terms aredetermined by the two boundary conditions u = 0 on the sphere <strong>and</strong> the <strong>for</strong>m ofψ <strong>for</strong> large r.21.9 Express cos 2 φ in terms of cos 2φ; T (ρ, φ) =A + B/2+(Bρ 2 /2a 2 )cos2φ.21.11 (A cos mx + B sin mx + C cosh mx + D sinh mx)cos(ωt + ɛ), with m 4 a 4 = ω 2 .21.13 E n =16ρA 2 c 2 /[(2n +1) 2 π 2 L]; E =2ρc 2 A 2 /L = ∫ A[2Tv/( 1 L)] dv.0 221.15 Note that the boundary value function is a square wave that is symmetric in φ.21.17 Since there is no heat flow at x = ±a, use a series of period 4a, u(x, 0) = 100 <strong>for</strong>0


PDES: SEPARATION OF VARIABLES AND OTHER METHODS21.25 The terms in G(r, r 0 ) that are additional to the fundamental solution are1∞∑ { [(x(−1) n − x0 ) 2 +(y − y 0 ) 2 +(z +(−1) n z 0 − nc) 2] −1/24πn=2+ [ (x − x 0 ) 2 +(y − y 0 ) 2 +(z +(−1) n z 0 + nc) 2] } −1/2.21.27 (a) As given in equation (21.86), but with r 0 replaced by r ′ .(b) Move the origin to r ′ <strong>and</strong> integrate the defining Green’s equation to obtain4πt 2 dG ∫ tdt − m2 G(t ′ )4πt ′2 dt ′ =1,0leading to G(t) =[−1/(4πt)]e −mt .(c) φ(r) =[−1/(4π)](p −1 e −mp − q −1 e −mq ), where p = |r − r 1 | <strong>and</strong> q = |r − r 2 | withr 1 =(x 1 ,y 1 ,z 1 )<strong>and</strong>r 2 =(−x 1 ,y 1 ,z 1 ).774


22Calculus of variationsIn chapters 2 <strong>and</strong> 5 we discussed how to find stationary values of functions of asingle variable f(x), of several variables f(x,y,...) <strong>and</strong> of constrained variables,where x,y,... are subject to the n constraints g i (x,y,...)=0,i =1, 2,...,n.Inallthese cases the <strong>for</strong>ms of the functions f <strong>and</strong> g i were known, <strong>and</strong> the problem wasone of finding the appropriate values of the variables x, y etc.We now turn to a different kind of problem in which we are interested inbringing about a particular condition <strong>for</strong> a given expression (usually maximisingor minimising it) by varying the functions on which the expression depends. Forinstance, we might want to know in what shape a fixed length of rope shouldbe arranged so as to enclose the largest possible area, or in what shape it willhang when suspended under gravity from two fixed points. In each case we areconcerned with a general maximisation or minimisation criterion by which thefunction y(x) that satisfies the given problem may be found.The calculus of variations provides a method <strong>for</strong> finding the function y(x).The problem must first be expressed in a mathematical <strong>for</strong>m, <strong>and</strong> the <strong>for</strong>mmost commonly applicable to such problems is an integral. Ineachoftheabovequestions, the quantity that has to be maximised or minimised by an appropriatechoice of the function y(x) may be expressed as an integral involving y(x) <strong>and</strong>the variables describing the geometry of the situation.In our example of the rope hanging from two fixed points, we need to findthe shape function y(x) that minimises the gravitational potential energy of therope. Each elementary piece of the rope has a gravitational potential energyproportional both to its vertical height above an arbitrary zero level <strong>and</strong> to thelength of the piece. There<strong>for</strong>e the total potential energy is given by an integral<strong>for</strong> the whole rope of such elementary contributions. The particular function y(x)<strong>for</strong> which the value of this integral is a minimum will give the shape assumed bythe hanging rope.775


CALCULUS OF VARIATIONSyabxFigure 22.1 Possible paths <strong>for</strong> the integral (22.1). The solid line is the curvealong which the integral is assumed stationary. The broken curves representsmall variations from this path.So in general we are led by this type of question to study the value of anintegral whose integr<strong>and</strong> has a specified <strong>for</strong>m in terms of a certain function<strong>and</strong> its derivatives, <strong>and</strong> to study how that value changes when the <strong>for</strong>m ofthe function is varied. Specifically, we aim to find the function that makes theintegral stationary, i.e. the function that makes the value of the integral a localmaximum or minimum. Note that, unless stated otherwise, y ′ is used to denotedy/dx throughout this chapter. We also assume that all the functions we need todeal with are sufficiently smooth <strong>and</strong> differentiable.Let us consider the integral22.1 The Euler–Lagrange equationI =∫ baF(y, y ′ ,x) dx, (22.1)where a, b <strong>and</strong> the <strong>for</strong>m of the function F are fixed by given considerations,e.g. the physics of the problem, but the curve y(x) is to be chosen so as tomake stationary the value of I, which is clearly a function, or more accurately afunctional, of this curve, i.e. I = I[y(x)]. Referring to figure 22.1, we wish to findthe function y(x) (given, say, by the solid line) such that first-order small changesin it (<strong>for</strong> example the two broken lines) will make only second-order changes inthe value of I.Writing this in a more mathematical <strong>for</strong>m, let us suppose that y(x) is thefunction required to make I stationary <strong>and</strong> consider making the replacementy(x) → y(x)+αη(x), (22.2)where the parameter α is small <strong>and</strong> η(x) is an arbitrary function with sufficientlyamenable mathematical properties. For the value of I to be stationary with respect776


22.2 SPECIAL CASESto these variations, we requiredIdα∣ = 0α=0<strong>for</strong> all η(x). (22.3)Substituting (22.2) into (22.1) <strong>and</strong> exp<strong>and</strong>ing as a Taylor series in α we obtainI(y, α) ==∫ ba∫ baF(y + αη, y ′ + αη ′ ,x) dxF(y, y ′ ,x) dx +∫ ba( )∂F ∂Fαη +∂y ∂y ′ αη′ dx +O(α 2 ).With this <strong>for</strong>m <strong>for</strong> I(y, α) the condition (22.3) implies that <strong>for</strong> all η(x) werequire∫ b( ∂FδI =a ∂y η + ∂F )∂y ′ η′ dx =0,where δI denotes the first-order variation in the value of I due to the variation(22.2) in the function y(x). Integrating the second term by parts this becomes[η ∂F ] b ∫ b[ ∂F∂y ′ +a a ∂y − d ( )] ∂Fdx ∂y ′ η(x) dx =0. (22.4)In order to simplify the result we will assume, <strong>for</strong> the moment, that the end-pointsare fixed, i.e. not only a <strong>and</strong> b are given but also y(a) <strong>and</strong>y(b). This restrictionmeans that we require η(a) =η(b) = 0, in which case the first term on the LHS of(22.4) equals zero at both end-points. Since (22.4) must be satisfied <strong>for</strong> arbitraryη(x), it is easy to see that we require∂F∂y = ddx( ∂F∂y ′ ). (22.5)This is known as the Euler–Lagrange (EL) equation, <strong>and</strong> is a differential equation<strong>for</strong> y(x), since the function F is known.22.2 Special casesIn certain special cases a first integral of the EL equation can be obtained <strong>for</strong> ageneral <strong>for</strong>m of F.22.2.1 F does not contain y explicitlyIn this case ∂F/∂y = 0, <strong>and</strong> (22.5) can be integrated immediately giving∂F∂y ′ = constant. (22.6)777


CALCULUS OF VARIATIONSB(b, y(b))dsdxdyA(a, y(a))Figure 22.2An arbitrary path between two fixed points.◮Show that the shortest curve joining two points is a straight line.Let the two points be labelled A <strong>and</strong> B <strong>and</strong> have coordinates (a, y(a)) <strong>and</strong> (b, y(b))respectively (see figure 22.2). Whatever the shape of the curve joining A to B, the lengthof an element of path ds is given byds = [ (dx) 2 +(dy) 2] 1/2=(1+y ′2 ) 1/2 dx,<strong>and</strong> hence the total path length along the curve is given byL =∫ ba(1 + y ′2 ) 1/2 dx. (22.7)We must now apply the results of the previous section to determine that path which makesL stationary (clearly a minimum in this case). Since the integral does not contain y (orindeed x) explicitly, we may use (22.6) to obtaink = ∂F y ′=∂y ′ (1 + y ′2 ) . 1/2where k is a constant. This is easily rearranged <strong>and</strong> integrated to giveky = x + c,(1 − k 2 )1/2which, as expected, is the equation of a straight line in the <strong>for</strong>m y = mx + c, withm = k/(1 − k 2 ) 1/2 . The value of m (or k) can be found by dem<strong>and</strong>ing that the straight linepasses through the points A <strong>and</strong> B <strong>and</strong> is given by m =[y(b) − y(a)]/(b − a). Substitutingthe equation of the straight line into (22.7) we find that, again as expected, the total pathlength is given byL 2 =[y(b) − y(a)] 2 +(b − a) 2 . ◭778


22.2 SPECIAL CASESydsdydxxFigure 22.3A convex closed curve that is symmetrical about the x-axis.22.2.2 F does not contain x explicitlyIn this case, multiplying the EL equation (22.5) by y ′ <strong>and</strong> using(dy ′ ∂F ) ( )dx ∂y ′ = y ′ d ∂F ′′ ∂Fdx ∂y ′ + y∂y ′we obtainy ′ ∂F ∂F+ y′′∂y ∂y ′ = d (y ′ ∂F )dx ∂y ′ .But since F is a function of y <strong>and</strong> y ′ only, <strong>and</strong> not explicitly of x, theLHSofthis equation is just the total derivative of F, namely dF/dx. Hence, integratingwe obtainF − y ′ ∂F∂y ′ = constant. (22.8)◮Find the closed convex curve of length l that encloses the greatest possible area.Without any loss of generality we can assume that the curve passes through the origin<strong>and</strong> can further suppose that it is symmetric with respect to the x-axis; this assumptionis not essential. Using the distance s along the curve, measured from the origin, as theindependent variable <strong>and</strong> y as the dependent one, we have the boundary conditionsy(0) = y(l/2) = 0. The element of area shown in figure 22.3 is then given by<strong>and</strong> the total area bydA = ydx= y [ (ds) 2 − (dy) 2] 1/2,A =2∫ l/20y(1 − y ′2 ) 1/2 ds; (22.9)here y ′ st<strong>and</strong>s <strong>for</strong> dy/ds rather than dy/dx. Since the integr<strong>and</strong> does not contain s explicitly,779


CALCULUS OF VARIATIONSwe can use (22.8) to obtain a first integral of the EL equation <strong>for</strong> y, namelyy(1 − y ′2 ) 1/2 + yy ′2 (1 − y ′2 ) −1/2 = k,where k is a constant. On rearranging this givesky ′ = ±(k 2 − y 2 ) 1/2 ,which, using y(0) = 0, integrates toy/k = sin(s/k). (22.10)The other end-point, y(l/2) = 0, fixes the value of k as l/(2π) to yieldy =l2π2πssin .lFrom this we obtain dy =cos(2πs/l) ds <strong>and</strong> since (ds) 2 =(dx) 2 +(dy) 2 we find also thatdx = ± sin(2πs/l) ds. This in turn can be integrated <strong>and</strong>, using x(0) = 0, gives x in termsof s asx − l2π = − l2π2πscos .lWe thus obtain the expected result that x <strong>and</strong> y lie on the circle of radius l/(2π) givenby(x − l ) 2+ y 2 = l22π 4π . 2Substituting the solution (22.10) into the expression <strong>for</strong> the total area (22.9), it is easilyverified that A = l 2 /(4π). A much quicker derivation of this result is possible using planepolar coordinates. ◭The previous two examples have been carried out in some detail, even thoughthe answers are more easily obtained in other ways, expressly so that the methodis transparent <strong>and</strong> the way in which it works can be filled in mentally at almostevery step. The next example, however, does not have such an intuitively obvioussolution.◮Two rings, each of radius a, are placed parallel with their centres 2b apart <strong>and</strong> on acommon normal. An open-ended axially symmetric soap film is <strong>for</strong>med between them (seefigure 22.4). Find the shape assumed by the film.Creating the soap film requires an energy γ per unit area (numerically equal to the surfacetension of the soap solution). So the stable shape of the soap film, i.e. the one thatminimises the energy, will also be the one that minimises the surface area (neglectinggravitational effects).It is obvious that any convex surface, shaped such as that shown as the broken line infigure 22.4(a), cannot be a minimum but it is not clear whether some shape intermediatebetween the cylinder shown by solid lines in (a), with area 4πab (or twice this <strong>for</strong> thedouble surface of the film), <strong>and</strong> the <strong>for</strong>m shown in (b), with area approximately 2πa 2 , willproduce a lower total area than both of these extremes. If there is such a shape (e.g. that infigure 22.4(c)), then it will be that which is the best compromise between two requirements,the need to minimise the ring-to-ring distance measured on the film surface (a) <strong>and</strong> theneed to minimise the average waist measurement of the surface (b).We take cylindrical polar coordinates as in figure 22.4(c) <strong>and</strong> let the radius of the soapfilm at height z be ρ(z) withρ(±b) =a. Counting only one side of the film, the element of780


22.3 SOME EXTENSIONSbzρ−ba(a) (b) (c)Figure 22.4Possible soap films between two parallel circular rings.surface area between z <strong>and</strong> z + dz isso the total surface area is given bydS =2πρ [ (dz) 2 +(dρ) 2] 1/2,S =2π∫ b−bρ(1 + ρ ′2 ) 1/2 dz. (22.11)Since the integr<strong>and</strong> does not contain z explicitly, we can use (22.8) to obtain an equation<strong>for</strong> ρ that minimises S, i.e.ρ(1 + ρ ′2 ) 1/2 − ρρ ′2 (1 + ρ ′2 ) −1/2 = k,where k is a constant. Multiplying through by (1 + ρ ′2 ) 1/2 , rearranging to find an explicitexpression <strong>for</strong> ρ ′ <strong>and</strong> integrating we findcosh −1 ρk = z k + c.where c is the constant of integration. Using the boundary conditions ρ(±b) =a, werequire c =0<strong>and</strong>k such that a/k =coshb/k (if b/a istoolarge,nosuchk can be found).Thus the curve that minimises the surface area isρ/k =cosh(z/k),<strong>and</strong> in profile the soap film is a catenary (see section 22.4) with the minimum distancefrom the axis equal to k. ◭22.3 Some extensionsIt is quite possible to relax many of the restrictions we have imposed so far. Forexample, we can allow end-points that are constrained to lie on given curves ratherthan being fixed, or we can consider problems with several dependent <strong>and</strong>/orindependent variables or higher-order derivatives of the dependent variable. Eachof these extensions is now discussed.781


CALCULUS OF VARIATIONS22.3.1 Several dependent variablesHere we have F = F(y 1 ,y 1 ′ ,y 2,y 2 ′ ,...,y n,y n,x)whereeachy ′ i = y i (x). The analysisin this case proceeds as be<strong>for</strong>e, leading to n separate but simultaneous equations<strong>for</strong> the y i (x),∂F∂y i= ddx( ∂F∂y ′ i), i =1, 2,...,n. (22.12)22.3.2 Several independent variablesWith n independent variables, we need to extremise multiple integrals of the <strong>for</strong>m∫ ∫ ∫ ()∂y ∂y ∂yI = ··· F y, , ,..., ,x 1 ,x 2 ,...,x n dx 1 dx 2 ···dx n .∂x 1 ∂x 2 ∂x nUsing the same kind of analysis as be<strong>for</strong>e, we find that the extremising functiony = y(x 1 ,x 2 ,...,x n ) must satisfy∂Fn∑( )∂y = ∂ ∂F, (22.13)∂x i ∂y xiwhere y xi st<strong>and</strong>s <strong>for</strong> ∂y/∂x i .i=122.3.3 Higher-order derivativesIf in (22.1) F = F(y, y ′ ,y ′′ ,...,y (n) ,x) then using the same method as be<strong>for</strong>e<strong>and</strong> per<strong>for</strong>ming repeated integration by parts, it can be shown that the requiredextremising function y(x) satisfies∂F∂y − ddx( ∂F∂y ′ )+ d2dx 2 ( ∂F∂y ′′ )− ···+(−1) n dndx n ( ∂F∂y (n) )=0, (22.14)provided that y = y ′ = ··· = y (n−1) = 0 at both end-points. If y, or any of itsderivatives, is not zero at the end-points then a corresponding contribution orcontributions will appear on the RHS of (22.14).22.3.4 Variable end-pointsWe now discuss the very important generalisation to variable end-points. Suppose,as be<strong>for</strong>e, we wish to find the function y(x) that extremises the integralI =∫ baF(y, y ′ ,x) dx,but this time we dem<strong>and</strong> only that the lower end-point is fixed, while we allowy(b) to be arbitrary. Repeating the analysis of section 22.1, we find from (22.4)782


22.3 SOME EXTENSIONSy(x)+η(x)∆yy(x)h(x, y) =0∆xbFigure 22.5 Variation of the end-point b along the curve h(x, y) =0.that we require[η ∂F ] b∂y ′ +a∫ ba[ ∂F∂y − d ( )] ∂Fdx ∂y ′ η(x) dx =0. (22.15)Obviously the EL equation (22.5) must still hold <strong>for</strong> the second term on the LHSto vanish. Also, since the lower end-point is fixed, i.e. η(a) = 0, the first term onthe LHS automatically vanishes at the lower limit. However, in order that it alsovanishes at the upper limit, we require in addition that∣∂F ∣∣∣x=b∂y ′ =0. (22.16)Clearly if both end-points may vary then ∂F/∂y ′ must vanish at both ends.An interesting <strong>and</strong> more general case is where the lower end-point is againfixed at x = a, but the upper end-point is free to lie anywhere on the curveh(x, y) = 0. Now in this case, the variation in the value of I due to the arbitraryvariation (22.2) is given to first order by[ ] ∂F b ∫ b( ∂FδI =∂y ′ η +a a ∂y − d )∂Fdx ∂y ′ ηdx+ F(b)∆x, (22.17)where ∆x is the displacement in the x-direction of the upper end-point, asindicated in figure 22.5, <strong>and</strong> F(b) is the value of F at x = b. In order <strong>for</strong> (22.17)to be valid, we of course require the displacement ∆x to be small.From the figure we see that ∆y = η(b) +y ′ (b)∆x. Since the upper end-pointmust lie on h(x, y) = 0 we also require that, at x = b,∂h ∂h∆x + ∆y =0,∂x ∂ywhich on substituting our expression <strong>for</strong> ∆y <strong>and</strong> rearranging becomes( )∂h ∂h+ y′ ∆x + ∂h η =0. (22.18)∂x ∂y ∂y783


CALCULUS OF VARIATIONSAx = x 0xyBFigure 22.6 A frictionless wire along which a small bead slides. We seek theshape of the wire that allows the bead to travel from the origin O to the linex = x 0 in the least possible time.Now, from (22.17) the condition δI = 0 requires, besides the EL equation, thatat x = b, the other two contributions cancel, i.e.F∆x + ∂F η =0. (22.19)∂y′Eliminating ∆x <strong>and</strong> η between (22.18) <strong>and</strong> (22.19) leads to the condition that atthe end-point(F − y ′ ∂F ) ∂h∂y ′ ∂y − ∂F ∂h∂y ′ =0. (22.20)∂xIn the special case where the end-point is free to lie anywhere on the vertical linex = b, we have ∂h/∂x = 1 <strong>and</strong> ∂h/∂y = 0. Substituting these values into (22.20),we recover the end-point condition given in (22.16).◮A frictionless wire in a vertical plane connects two points A <strong>and</strong> B, A being higher than B.Let the position of A be fixed at the origin of an xy-coordinate system, but allow B to lieanywhere on the vertical line x = x 0 (see figure 22.6). Find the shape of the wire such thatabeadplacedonitatA will slide under gravity to B in the shortest possible time.This is a variant of the famous brachistochrone (shortest time) problem, which is oftenused to illustrate the calculus of variations. Conservation of energy tells us that the particlespeed is given byv = dsdt = √ 2gy,where s is the path length along the wire <strong>and</strong> g is the acceleration due to gravity. Sincethe element of path length is ds =(1+y ′2 ) 1/2 dx, the total time taken to travel to the linex = x 0 is given by√∫ x=x0dst =x=0 v = √ 1 ∫ x01+y ′2dx.2g 0 yBecause the integr<strong>and</strong> does not contain x explicitly, we can use (22.8) with the specific<strong>for</strong>m F = √ 1+y ′2 / √ y to find a first integral; on simplification this yields[1/2y(1 + y )] ′2 = k,784


22.4 CONSTRAINED VARIATIONwhere k is a constant. Letting a = k 2 <strong>and</strong> solving <strong>for</strong> y ′ we findy ′ = dy √ a − ydx = ,ywhich on substituting y = a sin 2 θ integrates to givex = a (2θ − sin 2θ)+c.2Thus the parametric equations of the curve are given byx = b(φ − sin φ)+c, y = b(1 − cos φ),where b = a/2 <strong>and</strong>φ =2θ; they define a cycloid, the curve traced out by a point onthe rim of a wheel of radius b rolling along the x-axis. We must now use the end-pointconditions to determine the constants b <strong>and</strong> c. Since the curve passes through the origin,we see immediately that c =0.Nowsincey(x 0 ) is arbitrary, i.e. the upper end-point canlie anywhere on the curve x = x 0 , the condition (22.20) reduces to (22.16), so that we alsorequire∣ ∂F ∣∣∣x=x0 y ′= √ =0,∂y ′ y(1 + y ′2 ) ∣x=x0which implies that y ′ =0atx = x 0 . In words, the tangent to the cycloid at B must beparallel to the x-axis; this requires πb = x 0 . ◭22.4 Constrained variationJust as the problem of finding thestationary values of a function f(x, y) subject tothe constraint g(x, y) = constant is solved by means of Lagrange’s undeterminedmultipliers (see chapter 5), so the corresponding problem in the calculus ofvariations is solved by an analogous method.Suppose that we wish to find the stationary values ofI =∫ bsubject to the constraint that the value ofJ =a∫ baF(y, y ′ ,x) dx,G(y, y ′ ,x) dxis held constant. Following the method of Lagrange undetermined multipliers letus define a new functionalK = I + λJ =∫ ba(F + λG) dx,<strong>and</strong> find its unconstrained stationary values. Repeating the analysis of section 22.1we find that we require∂F∂y − d ( ) [ ∂F ∂Gdx ∂y ′ + λ∂y − d ( )] ∂Gdx ∂y ′ =0,785


CALCULUS OF VARIATIONS−ayOaxFigure 22.7A uni<strong>for</strong>m rope with fixed end-points suspended under gravity.which, together with the original constraint J = constant, will yield the requiredsolution y(x).This method is easily generalised to cases with more than one constraint by theintroduction of more Lagrange multipliers. If we wish to find the stationary valuesof an integral I subject to the multiple constraints that the values of the integralsJ i be held constant <strong>for</strong> i =1, 2,...,n, then we simply find the unconstrainedstationary values of the new integraln∑K = I + λ i J i .1◮Find the shape assumed by a uni<strong>for</strong>m rope when suspended by its ends from two pointsat equal heights.We will solve this problem using x (see figure 22.7) as the independent variable. Lettheropeoflength2L be suspended between the points x = ±a, y =0(L>a)<strong>and</strong>have uni<strong>for</strong>m linear density ρ. We then need to find the stationary value of the rope’sgravitational potential energy,∫∫ aI = −ρg yds= −ρg y(1 + y ′2 ) 1/2 dx,with respect to small changes in the <strong>for</strong>m of the rope but subject to the constraint thatthe total length of the rope remains constant, i.e.∫ ∫ aJ = ds = (1 + y ′2 ) 1/2 dx =2L.We thus define a new integral (omitting the factor −1 fromI <strong>for</strong> brevity)K = I + λJ =−a∫ a−a−a(ρgy + λ)(1 + y ′2 ) 1/2 dx<strong>and</strong> find its stationary values. Since the integr<strong>and</strong> does not contain the independentvariable x explicitly, we can use (22.8) to find the first integral:(ρgy + λ)(1+y ′2) 1/2 (− (ρgy + λ) 1+y ′2) −1/2y ′2 = k,786


22.5 PHYSICAL VARIATIONAL PRINCIPLESwhere k is a constant; this reduces to( ) 2 ρgy + λy ′2 =− 1.kMaking the substitution ρgy + λ = k cosh z, this can be integrated easily to give( )k ρgy + λρg cosh−1 = x + c,kwhere c is the constant of integration.We now have three unknowns, λ, k <strong>and</strong> c, that must be evaluated using the two endconditions y(±a) = 0 <strong>and</strong> the constraint J =2L. The end conditions giveρg(a + c)cosh = λ k k+ c)=coshρg(−a ,k<strong>and</strong> since a ≠ 0, these imply c =0<strong>and</strong>λ/k =cosh(ρga/k). Putting c =0intotheconstraint, in which y ′ = sinh(ρgx/k), we obtain∫ a[ ( ρgx)] 1/22L = 1+sinh 2 dx−ak= 2k ( ρga)ρg sinh .kCollecting together the values <strong>for</strong> the constants, the <strong>for</strong>m adopted by the rope is there<strong>for</strong>ey(x) = kρg[cosh( ρgxk)− cosh( ρgakwhere k is the solution of sinh(ρga/k) =ρgL/k. This curve is known as a catenary. ◭)],22.5 Physical variational principlesMany results in both classical <strong>and</strong> quantum physics can be expressed as variationalprinciples, <strong>and</strong> it is often when expressed in this <strong>for</strong>m that their physicalmeaning is most clearly understood. Moreover, once a physical phenomenon hasbeen written as a variational principle, we can use all the results derived in thischapter to investigate its behaviour. It is usually possible to identify conservedquantities, or symmetries of the system of interest, that otherwise might be foundonly with considerable ef<strong>for</strong>t. From the wide range of physical variational principleswe will select two examples from familiar areas of classical physics, namelygeometric optics <strong>and</strong> mechanics.22.5.1 Fermat’s principle in opticsFermat’s principle in geometrical optics states that a ray of light travelling in aregion of variable refractive index follows a path such that the total optical pathlength (physical length × refractive index) is stationary.787


CALCULUS OF VARIATIONSyBn 2xθ 1θ 2n 1AFigure 22.8 Path of a light ray at the plane interface between media withrefractive indices n 1 <strong>and</strong> n 2 ,wheren 2


22.5 PHYSICAL VARIATIONAL PRINCIPLESyOdx l xFigure 22.9 Transverse displacement on a taut string that is fixed at twopoints a distance l apart.states that in moving from one configuration at time t 0 to another at time t 1 themotion of such a system is such as to make∫ t1L = L(q 1 ,q 2 ...,q n , ˙q 1 , ˙q 2 ,...,˙q n ,t) dt (22.21)t 0stationary. The Lagrangian L is defined, in terms of the kinetic energy T <strong>and</strong>the potential energy V (with respect to some reference situation), by L = T − V .Here V is a function of the q i only, not of the ˙q i . Applying the EL equation to Lwe obtain Lagrange’s equations,∂L= d ( ) ∂L, i =1, 2,...,n.∂q i dt ∂˙q i◮Using Hamilton’s principle derive the wave equation <strong>for</strong> small transverse oscillations of ataut string.In this example we are in fact considering a generalisation of (22.21) to a case involvingone isolated independent coordinate t, togetherwithacontinuum in which the q i becomethe continuous variable x. The expressions <strong>for</strong> T <strong>and</strong> V there<strong>for</strong>e become integrals over xrather than sums over the label i.If ρ <strong>and</strong> τ are the local density <strong>and</strong> tension of the string, both of which may depend onx, then, referring to figure 22.9, the kinetic <strong>and</strong> potential energies of the string are givenby<strong>and</strong> (22.21) becomes∫ lT =0L = 1 2ρ2∫ t1( ) 2 ∫ ∂ylτdx, V =∂t0 2t 0dt∫ l0[ρ( ) 2 ∂y− τ∂t789( ) 2 ∂ydx∂x( ) ] 2 ∂ydx.∂x


CALCULUS OF VARIATIONSUsing (22.13) <strong>and</strong> the fact that y does not appear explicitly, we obtain(∂ρ ∂y )− ∂ (τ ∂y )=0.∂t ∂t ∂x ∂xIf, in addition, ρ <strong>and</strong> τ do not depend on x or t then∂ 2 y∂x = 1 ∂ 2 y2 c 2 ∂t , 2where c 2 = τ/ρ. This is the wave equation <strong>for</strong> small transverse oscillations of a tautuni<strong>for</strong>m string. ◭22.6 General eigenvalue problemsWe have seen in this chapter that the problem of finding a curve that makes thevalue of a given integral stationary when the integral is taken along the curveresults, in each case, in a differential equation <strong>for</strong> the curve. It is not a greatextension to ask whether this may be used to solve differential equations, bysetting up a suitable variational problem <strong>and</strong> then seeking ways other than theEuler equation of finding or estimating stationary solutions.We shall be concerned with differential equations of the <strong>for</strong>m Ly = λρ(x)y,where the differential operator L is self-adjoint, so that L = L † (with appropriateboundary conditions on the solution y) <strong>and</strong>ρ(x) is some weight function, asdiscussed in chapter 17. In particular, we will concentrate on the Sturm–Liouvilleequation as an explicit example, but much of what follows can be applied toother equations of this type.We have already discussed the solution of equations of the Sturm–Liouvilletype in chapter 17 <strong>and</strong> the same notation will be used here. In this section,however, we will adopt a variational approach to estimating the eigenvalues ofsuch equations.Suppose we search <strong>for</strong> stationary values of the integral∫ b []I = p(x)y ′2 (x) − q(x)y 2 (x) dx, (22.22)awith y(a) =y(b) =0<strong>and</strong>p <strong>and</strong> q any sufficiently smooth <strong>and</strong> differentiablefunctions of x. However, in addition we impose a normalisation conditionJ =∫ baρ(x)y 2 (x) dx = constant. (22.23)Here ρ(x) is a positive weight function defined in the interval a ≤ x ≤ b, butwhich may in particular cases be a constant.Then, as in section 22.4, we use undetermined Lagrange multipliers, § <strong>and</strong>§ We use −λ, rather than λ, so that the final equation (22.24) appears in the conventional Sturm–Liouville <strong>for</strong>m.790


22.6 GENERAL EIGENVALUE PROBLEMSconsider K = I − λJ given byK =∫ ba[py ′2 − (q + λρ)y 2] dx.On application of the EL equation (22.5) this yields(dp dy )+ qy + λρy =0, (22.24)dx dxwhich is exactly the Sturm–Liouville equation (17.34), with eigenvalue λ. Now,since both I <strong>and</strong> J are quadratic in y <strong>and</strong> its derivative, finding stationary valuesof K is equivalent to finding stationary values of I/J. This may also be shownby considering the functional Λ = I/J, <strong>for</strong> whichδΛ =(δI/J) − (I/J 2 ) δJ=(δI − ΛδJ)/J= δK/J.Hence, extremising Λ is equivalent to extremising K. Thus we have the importantresult that finding functions y that make I/J stationary is equivalent to findingfunctions y that are solutions of the Sturm–Liouville equation; the resulting valueof I/J equals the corresponding eigenvalue of the equation.Of course this does not tell us how to find such a function y <strong>and</strong>, naturally, tohave to do this by solving (22.24) directly defeats the purpose of the exercise. Wewill see in the next section how some progress can be made. It is worth recallingthat the functions p(x), q(x) <strong>and</strong>ρ(x) can have many different <strong>for</strong>ms, <strong>and</strong> so(22.24) represents quite a wide variety of equations.We now recall some properties of the solutions of the Sturm–Liouville equation.The eigenvalues λ i of (22.24) are real <strong>and</strong> will be assumed non-degenerate (<strong>for</strong>simplicity). We also assume that the corresponding eigenfunctions have been madereal, so that normalised eigenfunctions y i (x) satisfy the orthogonality relation (asin (17.24))∫ bay i y j ρdx= δ ij . (22.25)Further, we take the boundary condition in the <strong>for</strong>m[ ] x=by i py j′ = 0; (22.26)x=athis can be satisfied by y(a) =y(b) = 0, but also by many other sets of boundaryconditions.791


CALCULUS OF VARIATIONS◮Show that∫ b(y′j py ′ i − y j qy i)dx = λi δ ij . (22.27)aLet y i be an eigenfunction of (22.24), corresponding to a particular eigenvalue λ i ,sothat( )py′ ′i +(q + λi ρ)y i =0.Multiplying this through by y j <strong>and</strong> integrating from a to b (the first term by parts) weobtain( )[y ] b ∫ b∫ bj py′i y j(py ′ i) ′ dx + y j (q + λ i ρ)y i dx =0. (22.28)−aaThe first term vanishes by virtue of (22.26), <strong>and</strong> on rearranging the other terms <strong>and</strong> using(22.25), we find the result (22.27). ◭We see at once that, if the function y(x) minimises I/J, i.e. satisfies the Sturm–Liouville equation, then putting y i = y j = y in (22.25) <strong>and</strong> (22.27) yields J <strong>and</strong>I respectively on the left-h<strong>and</strong> sides; thus, as mentioned above, the minimisedvalue of I/J is just the eigenvalue λ, introduced originally as the undeterminedmultiplier.a◮For a function y satisfying the Sturm–Liouville equation verify that, provided (22.26) issatisfied, λ = I/J.Firstly, we multiply (22.24) through by y to givey(py ′ ) ′ + qy 2 + λρy 2 =0.Now integrating this expression by parts we have[ ] b ∫ b) ∫ bypy ′ −(py ′2 − qy 2 dx + λ ρy 2 dx =0.aaThe first term on the LHS is zero, the second is simply −I <strong>and</strong> the third is λJ. Thusλ = I/J. ◭a22.7 Estimation of eigenvalues <strong>and</strong> eigenfunctionsSince the eigenvalues λ i of the Sturm–Liouville equation are the stationary valuesof I/J (see above), it follows that any evaluation of I/J must yield a value that liesbetween the lowest <strong>and</strong> highest eigenvalues of the corresponding Sturm–Liouvilleequation, i.e.λ min ≤ I J ≤ λ max,where, depending on the equation under consideration, either λ min = −∞ <strong>and</strong>792


22.7 ESTIMATION OF EIGENVALUES AND EIGENFUNCTIONSλ max is finite, or λ max = ∞ <strong>and</strong> λ min is finite. Notice that here we have departedfrom direct consideration of the minimising problem <strong>and</strong> made a statement abouta calculation in which no actual minimisation is necessary.Thus, as an example, <strong>for</strong> an equation with a finite lowest eigenvalue λ 0 anyevaluation of I/J provides an upper bound on λ 0 . Further, we will now show thatthe estimate λ obtained is a better estimate of λ 0 than the estimated (guessed)function y is of y 0 , the true eigenfunction corresponding to λ 0 . The sense in which‘better’ is used here will be clear from the final result.Firstly, we exp<strong>and</strong> the estimated or trial function y in terms of the completeset y i :y = y 0 + c 1 y 1 + c 2 y 2 + ··· ,where, if a good trial function has been guessed, the c i will be small. Using (22.25)we have immediately that J =1+ ∑ i |c i| 2 . The other required integral isI =∫ ba[ (p y 0 ′ + ∑ i2 (c i y i) ′ − q y 0 + ∑ i) ] 2c i y i dx.On multiplying out the squared terms, all the cross terms vanish because of(22.27) to leaveλ = I J= λ 0 + ∑ i |c i| 2 λ i1+ ∑ j |c j| 2= λ 0 + ∑ i|c i | 2 (λ i − λ 0 )+O(c 4 ).Hence λ differs from λ 0 by a term second order in the c i , even though y differedfrom y 0 by a term first order in the c i ; this is what we aimed to show. We noticeincidentally that, since λ 0


CALCULUS OF VARIATIONSy(x)1(c)0.80.6(b)(a)0.4(d)0.20.20.40.60.8x1Figure 22.10 Trial solutions used to estimate the lowest eigenvalue λ of−y ′′ = λy with y(0) = y ′ (1) = 0. They are: (a) y = sin(πx/2), the exact result;(b) y =2x − x 2 ;(c)y = x 3 − 3x 2 +3x; (d)y =sin 2 (πx/2).◮Estimate the lowest eigenvalue of the equationwith boundary conditions− d2 y= λy, 0 ≤ x ≤ 1, (22.29)dx2 y(0) = 0, y ′ (1) = 0. (22.30)We need to find the lowest value λ 0 of λ <strong>for</strong> which (22.29) has a solution y(x) that satisfies(22.30). The exact answer is of course y = A sin(xπ/2) <strong>and</strong> λ 0 = π 2 /4 ≈ 2.47.Firstly we note that the Sturm–Liouville equation reduces to (22.29) if we take p(x) =1,q(x) =0<strong>and</strong>ρ(x) = 1 <strong>and</strong> that the boundary conditions satisfy (22.26). Thus we are ableto apply the previous theory.We will use three trial functions so that the effect on the estimate of λ 0 of making betteror worse ‘guesses’ can be seen. One further preliminary remark is relevant, namely that theestimate is independent of any constant multiplicative factor in the function used. Thisis easily verified by looking at the <strong>for</strong>m of I/J. We normalise each trial function so thaty(1) = 1, purely in order to facilitate comparison of the various function shapes.Figure 22.10 illustrates the trial functions used, curve (a) being the exact solutiony =sin(πx/2). The other curves are (b) y(x) =2x − x 2 ,(c)y(x) =x 3 − 3x 2 +3x, <strong>and</strong>(d)y(x) =sin 2 (πx/2). The choice of trial function is governed by the following considerations:(i) the boundary conditions (22.30) must be satisfied.(ii) a ‘good’ trial function ought to mimic the correct solution as far as possible, butit may not be easy to guess even the general shape of the correct solution in somecases.(iii) the evaluation of I/J should be as simple as possible.794


22.8 ADJUSTMENT OF PARAMETERSIt is easily verified that functions (b), (c) <strong>and</strong> (d) all satisfy (22.30) but, so far as mimickingthe correct solution is concerned, we would expect from the figure that (b) would besuperior to the other two. The three evaluations are straight<strong>for</strong>ward, using (22.22) <strong>and</strong>(22.23):∫ 10λ b =(2 − 2x)2 dx∫ 1(2x − 0 x2 ) 2 dx = 4/38/15 =2.50∫ 10λ c =(3x2 − 6x +3) 2 dx∫ 10 (x3 − 3x 2 +3x) 2 dx = 9/59/14 =2.80∫ 10λ d =(π2 /4) sin 2 (πx) dx∫ 1= π2 /80 sin4 (πx/2) dx 3/8 =3.29.We expected all evaluations to yield estimates greater than the lowest eigenvalue, 2.47,<strong>and</strong> this is indeed so. From these trials alone we are able to say (only) that λ 0 ≤ 2.50.As expected, the best approximation (b) to the true eigenfunction yields the lowest, <strong>and</strong>there<strong>for</strong>e the best, upper bound on λ 0 . ◭We may generalise the work of this section to other differential equations ofthe <strong>for</strong>m Ly = λρy, whereL = L † . In particular, one findswhere I <strong>and</strong> J are now given byI =∫ baλ min ≤ I J ≤ λ max,y ∗ (Ly) dx <strong>and</strong> J =∫ baρy ∗ ydx. (22.31)It is straight<strong>for</strong>ward to show that, <strong>for</strong> the special case of the Sturm–Liouvilleequation, <strong>for</strong> whichLy = −(py ′ ) ′ − qy,the expression <strong>for</strong> I in (22.31) leads to (22.22).22.8 Adjustment of parametersInstead of trying to estimate λ 0 by selecting a large number of different trialfunctions, we may also use trial functions that include one or more parameterswhich themselves may be adjusted to give the lowest value to λ = I/J <strong>and</strong>hence the best estimate of λ 0 . The justification <strong>for</strong> this method comes from theknowledge that no matter what <strong>for</strong>m of function is chosen, nor what values areassigned to the parameters, provided the boundary conditions are satisfied λ cannever be less than the required λ 0 .To illustrate this method an example from quantum mechanics will be used.The time-independent Schrödinger equation is <strong>for</strong>mally written as the eigenvalueequation Hψ = Eψ, whereH is a linear operator, ψ the wavefunction describinga quantum mechanical system <strong>and</strong> E the energy of the system. The energy795


CALCULUS OF VARIATIONSoperator H is called the Hamiltonian <strong>and</strong> <strong>for</strong> a particle of mass m moving in aone-dimensional harmonic oscillator potential is given byH = − 22m dx 2 + kx22 , (22.32)where is Planck’s constant divided by 2π.◮Estimate the ground-state energy of a quantum harmonic oscillator.Using (22.32) in Hψ = Eψ, the Schrödinger equation is− 2 d 2 ψ2m dx + kx2 ψ = Eψ, −∞


22.9 EXERCISES22.9 Exercises22.1 A surface of revolution, whose equation in cylindrical polar coordinates is ρ =ρ(z), is bounded by the circles ρ = a, z = ±c (a >c). Show that the functionthat makes the surface integral I = ∫ ρ −1/2 dS stationary with respect to smallvariationsisgivenbyρ(z) =k + z 2 /(4k), where k =[a ± (a 2 − c 2 ) 1/2 ]/2.22.2 Show that the lowest value of the integral∫ B(1 + y ′2 ) 1/2dx,A ywhere A is (−1, 1) <strong>and</strong> B is (1, 1), is 2 ln(1+ √ 2). Assume that the Euler–Lagrangeequation gives a minimising curve.22.3 The refractive index n of a medium is a function only of the distance r from afixed point O. Prove that the equation of a light ray, assumed to lie in a planethrough O, travelling in the medium satisfies (in plane polar coordinates)( ) 21 dr= r2 n 2 (r)r 2 dφ a 2 n 2 (a) − 1,where a is the distance of the ray from O at the point at which dr/dφ =0.If n = [1+(α 2 /r 2 )] 1/2 <strong>and</strong> the ray starts <strong>and</strong> ends far from O, find its deviation(the angle through which the ray is turned), if its minimum distance from O is a.22.4 The Lagrangian <strong>for</strong> a π-meson is given byL(x,t)= 1 (˙φ 2 −|∇φ| 2 − µ 2 φ 2 ),2where µ is the meson mass <strong>and</strong> φ(x,t) is its wavefunction. Assuming Hamilton’sprinciple, find the wave equation satisfied by φ.22.5 Prove the following results about general systems.(a) For a system described in terms of coordinates q i <strong>and</strong> t, show that if t doesnot appear explicitly in the expressions <strong>for</strong> x, y <strong>and</strong> z (x = x(q i ,t), etc.) thenthe kinetic energy T is a homogeneous quadratic function of the ˙q i (it mayalso involve the q i ). Deduce that ∑ i ˙q i(∂T/∂˙q i )=2T .(b) Assuming that the <strong>for</strong>ces acting on the system are derivable from a potentialV , show, by expressing dT /dt in terms of q i <strong>and</strong> ˙q i ,thatd(T + V )/dt =0.22.6 For a system specified by the coordinates q <strong>and</strong> t, show that the equation ofmotion is unchanged if the Lagrangian L(q, ˙q, t) is replaced byL 1 = L + dφ(q,t) ,dtwhere φ is an arbitrary function. Deduce that the equation of motion of a particlethat moves in one dimension subject to a <strong>for</strong>ce −dV (x)/dx (x being measuredfrom a point O) is unchanged if O is <strong>for</strong>ced to move with a constant velocity v(x still being measured from O).22.7 In cylindrical polar coordinates, the curve (ρ(θ),θ,αρ(θ)) lies on the surface ofthe cone z = αρ. Show that geodesics (curves of minimum length joining twopoints) on the cone satisfyρ 4 = c 2 [β 2 ρ ′2 + ρ 2 ],where c is an arbitrary constant, but β has to have a particular value. Determinethe <strong>for</strong>m of ρ(θ) <strong>and</strong> hence find the equation of the shortest path on the conebetween the points (R,−θ 0 ,αR)<strong>and</strong>(R,θ 0 ,αR).[ You will find it useful to determine the <strong>for</strong>m of the derivative of cos −1 (u −1 ). ]797


CALCULUS OF VARIATIONS22.8 Derive the differential equations <strong>for</strong> the plane-polar coordinates, r <strong>and</strong> φ, ofaparticle of unit mass moving in a field of potential V (r). Find the <strong>for</strong>m of V ifthe path of the particle is given by r = a sin φ.22.9 You are provided with a line of length πa/2 <strong>and</strong> negligible mass <strong>and</strong> some leadshot of total mass M. Use a variational method to determine how the lead shotmust be distributed along the line if the loaded line is to hang in a circular arcof radius a when its ends are attached to two points at the same height. Measurethe distance s along the line from its centre.22.10 Extend the result of subsection 22.2.2 to the case of several dependent variablesy i (x), showing that, if x does not appear explicitly in the integr<strong>and</strong>, then a firstintegral of the Euler–Lagrange equations isn∑F − y ′ ∂Fi =constant.∂y ′ i=1 i22.11 A general result is that light travels through a variable medium by a path whichminimises the travel time (this is an alternative <strong>for</strong>mulation of Fermat’s principle).With respect to a particular cylindrical polar coordinate system (ρ, φ, z), the speedof light v(ρ, φ) is independent of z. If the path of the light is parameterised asρ = ρ(z),φ= φ(z), use the result of the previous exercise to show thatv 2 (ρ ′2 + ρ 2 φ ′2 +1)is constant along the path.For the particular case when v = v(ρ) =b(a 2 +ρ 2 ) 1/2 , show that the two Euler–Lagrange equations have a common solution in which the light travels along ahelical path given by φ = Az + B, ρ = C, provided that A has a particular value.22.12 Light travels in the vertical xz-plane through a slab of material which lies betweenthe planes z = z 0 <strong>and</strong> z =2z 0 , <strong>and</strong> in which the speed of light v(z) =c 0 z/z 0 .Using the alternative <strong>for</strong>mulation of Fermat’s principle, given in the previousquestion, show that the ray paths are arcs of circles.Deduce that, if a ray enters the material at (0,z 0 ) at an angle to the vertical,π/2 − θ, ofmorethan30 ◦ , then it does not reach the far side of the slab.22.13 A dam of capacity V (less than πb 2 h/2) is to be constructed on level ground nextto a long straight wall which runs from (−b, 0) to (b, 0).Thisistobeachievedbyjoining the ends of a new wall, of height h, to those of the existing wall. Showthat, in order to minimise the length L of new wall to be built, it should <strong>for</strong>mpart of a circle, <strong>and</strong> that L is then given by∫ bdx−b (1 − λ 2 x 2 ) , 1/2where λ is found fromVhb = sin−1 µ− (1 − µ2 ) 1/22 µ 2 µ<strong>and</strong> µ = λb.22.14 In the brachistochrone problem of subsection 22.3.4 show that if the upper endpointcan lie anywhere on the curve h(x, y) = 0, then the curve of quickest descenty(x) meetsh(x, y) = 0 at right angles.22.15 The Schwarzchild metric <strong>for</strong> the static field of a non-rotating spherically symmetricblack hole of mass M is given by((ds) 2 = c 2 1 − 2GM )(dt) 2 −c 2 r(dr) 21 − 2GM/(c 2 r) − r2 (dθ) 2 − r 2 sin 2 θ (dφ) 2 .Considering only motion confined to the plane θ = π/2, <strong>and</strong> assuming that the798


22.9 EXERCISESpath of a small test particle is such as to make ∫ ds stationary, find two firstintegrals of the equations of motion. From their Newtonian limits, in whichGM/r, ṙ 2 <strong>and</strong> r 2 ˙φ 2 are all ≪ c 2 , identify the constants of integration.22.16 Use result (22.27) to evaluateJ =∫ 1−1(1 − x 2 )P ′ m(x)P ′ n(x) dx,where P m (x) is a Legendre polynomial of order m.22.17 Determine the minimum value that the integralJ =∫ 10[x 4 (y ′′ ) 2 +4x 2 (y ′ ) 2 ] dxcan have, given that y is not singular at x =0<strong>and</strong>thaty(1) = y ′ (1) = 1. Assumethat the Euler–Lagrange equation gives the lower limit, <strong>and</strong> verify retrospectivelythat your solution makes the first term on the LHS of equation (22.15) vanish.22.18 Show that y ′′ − xy + λx 2 y = 0 has a solution <strong>for</strong> which y(0) = y(1) = 0 <strong>and</strong>λ ≤ 147/4.22.19 Find an appropriate, but simple, trial function <strong>and</strong> use it to estimate the lowesteigenvalue λ 0 of Stokes’ equation,d 2 y+ λxy =0, with y(0) = y(π) =0.dx2 Explain why your estimate must be strictly greater than λ 0 .22.20 Estimate the lowest eigenvalue, λ 0 , of the equationd 2 ydx − 2 x2 y + λy =0, y(−1) = y(1) = 0,using a quadratic trial function.22.21 A drumskin is stretched across a fixed circular rim of radius a. Small transversevibrations of the skin have an amplitude z(ρ, φ, t) thatsatisfies∇ 2 z = 1 ∂ 2 zc 2 ∂t 2in plane polar coordinates. For a normal mode independent of azimuth, z =Z(ρ)cosωt, find the differential equation satisfied by Z(ρ). By using a trialfunction of the <strong>for</strong>m a ν − ρ ν , with adjustable parameter ν, obtain an estimate <strong>for</strong>the lowest normal mode frequency.[ The exact answer is (5.78) 1/2 c/a.]22.22 Consider the problem of finding the lowest eigenvalue, λ 0 , of the equation(1 + x 2 ) d2 y dy+2x + λy =0, y(±1) = 0.dx2 dx(a) Recast the problem in variational <strong>for</strong>m, <strong>and</strong> derive an approximation λ 1 toλ 0 by using the trial function y 1 (x) =1− x 2 .(b) Show that an improved estimate λ 2 is obtained by using y 2 (x) =cos(πx/2).(c) Prove that the estimate λ(γ) obtained by taking y 1 (x) +γy 2 (x) asthetrialfunction isλ(γ) = 64/15 + 64γ/π − 384γ/π3 +(π 2 /3+1/2)γ 2.16/15 + 64γ/π 3 + γ 2Investigate λ(γ) numerically as γ is varied, or, more simply, show thatλ(−1.80) = 3.668, an improvement on both λ 1 <strong>and</strong> λ 2 .799


CALCULUS OF VARIATIONS22.23 For the boundary conditions given below, obtain a functional Λ(y) whose stationaryvalues give the eigenvalues of the equation(1 + x) d2 y dy+(2+x)dx2 dx + λy =0, y(0) = 0, y′ (2) = 0.Derive an approximation to the lowest eigenvalue λ 0 using the trial functiony(x) =xe −x/2 . For what value(s) of γ wouldy(x) =xe −x/2 + β sin γxbe a suitable trial function <strong>for</strong> attempting to obtain an improved estimate of λ 0 ?22.24 This is an alternative approach to the example in section 22.8. Using the notation∫of that section, the expectation value of the energy of the state ψ is given byψ ∗ Hψdv. Denote the eigenfunctions of H by ψ i ,sothatHψ i = E i ψ i , <strong>and</strong>, sinceH is self-adjoint (Hermitian), ∫ ψj ∗ψ i dv = δ ij .(a) By writing any function ψ as ∑ c j ψ j <strong>and</strong> following an argument similar tothat in section 22.7, show that∫ψ ∗ HψdvE = ∫ ≥ E 0 ,ψ∗ ψdvthe energy of the lowest state. This is the Rayleigh–Ritz principle.(b) Using the same trial function as in section 22.8, ψ =exp(−αx 2 ), show thatthe same result is obtained.22.25 This is an extension to section 22.8 <strong>and</strong> the previous question. With the groundstate(i.e. the lowest-energy) wavefunction as exp(−αx 2 ), take as a trial functionthe orthogonal wave function x 2n+1 exp(−αx 2 ), using the integer n as a variableparameter. Use either Sturm–Liouville theory or the Rayleigh–Ritz principle toshow that the energy of the second lowest state of a quantum harmonic oscillatoris ≤ 3ω/2.22.26 The Hamiltonian H <strong>for</strong> the hydrogen atom is− 22m ∇2 −q24πɛ 0 r .For a spherically symmetric state, as may be assumed <strong>for</strong> the ground state, theonly relevant part of ∇ 2 is that involving differentiation with respect to r.(a) Define the integrals J n byJ n =∫ ∞0r n e −2βr dr∫<strong>and</strong> show that, <strong>for</strong> a trial wavefunction of the <strong>for</strong>m exp(−βr) withβ>0,ψ ∗ Hψdv <strong>and</strong> ∫ ψ ∗ ψdv(see exercise 22.24(a)) can be expressed as aJ 1 − bJ 2<strong>and</strong> cJ 2 respectively, where a, b <strong>and</strong> c are factors which you should determine.(b) Show that the estimate of E is minimised when β = mq 2 /(4πɛ 0 2 ).(c) Hence find an upper limit <strong>for</strong> the ground-state energy of the hydrogen atom.In fact, exp(−βr) is the correct <strong>for</strong>m <strong>for</strong> the wavefunction <strong>and</strong> the limit givesthe actual value.22.27 The upper <strong>and</strong> lower surfaces of a film of liquid, which has surface energy perunit area (surface tension) γ <strong>and</strong> density ρ, have equations z = p(x) <strong>and</strong>z = q(x),respectively. The film has a given volume V (per unit depth in the y-direction)<strong>and</strong> lies in the region −L


22.10 HINTS AND ANSWERStotal energy (per unit depth) of the film consists of its surface energy <strong>and</strong> itsgravitational energy, <strong>and</strong> is expressed by∫ L∫ L]E = ρg (p 2 − q 2 ) dx + γ[(1 + p ′2 ) 1/2 +(1+q ′2 ) 1/2 dx.2−L(a) Express V in terms of p <strong>and</strong> q.(b) Show that, if the total energy is minimised, p <strong>and</strong> q must satisfyp ′2(1 + p ′2 ) − q ′2=constant.1/2 (1 + q ′2 )1/2(c) As an approximate solution, consider the equationsp = a(L −|x|), q = b(L −|x|),where a <strong>and</strong> b are sufficiently small that a 3 <strong>and</strong> b 3 can be neglected comparedwith unity. Find the values of a <strong>and</strong> b that minimise E.22.28 A particle of mass m moves in a one-dimensional potential well of the <strong>for</strong>mV (x) =−µ 2 α 2m sech 2 αx,where µ <strong>and</strong> α are positive constants. As in exercise 22.26, the expectation value〈E〉 of the energy of the system is ∫ ψ ∗ Hψdx, where the self-adjoint operatorH is given by −( 2 /2m)d 2 /dx 2 + V (x). Using trial wavefunctions of the <strong>for</strong>my = A sech βx, show the following:(a) <strong>for</strong> µ = 1, there is an exact eigenfunction of H, with a corresponding 〈E〉 ofhalf of the maximum depth of the well;(b) <strong>for</strong> µ = 6, the ‘binding energy’ of the ground state is at least 10 2 α 2 /(3m).[ You will find it useful to note that <strong>for</strong> u, v ≥ 0, sech u sech v ≥ sech(u + v). ]22.29 The Sturm–Liouville equation can be extended to two independent variables, x<strong>and</strong> z, with little modification. In equation (22.22), y ′2 is replaced by (∇y) 2 <strong>and</strong>the integrals of the various functions of y(x, z) become two-dimensional, i.e. theinfinitesimal is dx dz.The vibrations of a trampoline 4 units long <strong>and</strong> 1 unit wide satisfy the equation∇ 2 y + k 2 y =0.By taking the simplest possible permissible polynomial as a trial function, showthat the lowest mode of vibration has k 2 ≤ 10.63 <strong>and</strong>, by direct solution, that theactual value is 10.49.−L22.10 Hints <strong>and</strong> answers22.1 Note that the integr<strong>and</strong>, 2πρ 1/2 (1 + ρ ′2 ) 1/2 , does not contain z explicitly.22.3 I = ∫ n(r)[r 2 +(dr/dφ) 2 ] 1/2 dφ. Take axes such that φ = 0 when r = ∞.If β =(π − deviation angle)/2 thenβ = φ at r = a, <strong>and</strong> the equation reduces to∫β∞(a 2 + α 2 ) = dr1/2 −∞ r(r 2 − a 2 ) , 1/2which can be evaluated by putting r = a(y + y −1 )/2, or successively r = a cosh ψ,y =expψ to yield a deviation of π[(a 2 + α 2 ) 1/2 − a]/a.801


CALCULUS OF VARIATIONS22.5 (a) ∂x/∂t =0<strong>and</strong>soẋ = ∑ i ˙q i∂x/∂q i ; (b) use∑( )d ∂T˙q i = d dt ∂˙q i dt (2T ) − ∑ ii22.7 Use result (22.8); β 2 =1+α 2 .¨q i∂T∂˙q i.Put ρ = uc to obtain dθ/du = β/[u(u 2 − 1) 1/2 ]. Remember that cos −1 is amultivalued function; ρ(θ) =[R cos(θ 0 /β)]/[cos(θ/β)].22.9 −λy ′ (1 − y ′2 ) −1/2 = 2gP(s), y = y(s), P (s) = ∫ s0 ρ(s′ ) ds ′ . The solution, y =−a cos(s/a), <strong>and</strong> 2P (πa/4) = M together give λ = −gM. Therequiredρ(s) isgiven by [M/(2a)] sec 2 (s/a).22.11 Note that the φ E–L equation is automatically satisfied if v ≠ v(φ). A =1/a.22.13 Circle is λ 2 x 2 +[λy +(1− λ 2 b 2 ) 1/2 ] 2 =1.Usethefactthat ∫ ydx = V/h todetermine the condition on λ.22.15 Denoting (ds) 2 /(dt) 2 by f 2 , the Euler–Lagrange equation <strong>for</strong> φ gives r 2 ˙φ = Af,where A corresponds to the angular momentum of the particle. Use the resultof exercise 22.10 to obtain c 2 − (2GM/r) =Bf, where, to first order in smallquantities,cB = c 2 − GM + 1 r 2 (ṙ2 + r 2 ˙φ2 ),which reads ‘total energy = rest mass + gravitational energy + radial <strong>and</strong>azimuthal kinetic energy’.22.17 Convert the equation to the usual <strong>for</strong>m, by writing y ′ (x) =u(x), <strong>and</strong> obtainx 2 u ′′ +4xu ′ − 4u = 0 with general solution Ax −4 + Bx. Integrating a second time<strong>and</strong> using the boundary conditions gives y(x) =(1+x 2 )/2 <strong>and</strong>J =1;η(1) = 0,since y ′ (1) is fixed, <strong>and</strong> ∂F/∂u ′ =2x 4 u ′ =0atx =0.22.19 Using y =sinx as a trial function shows that λ 0 ≤ 2/π. The estimate must be>λ 0 since the trial function does not satisfy the original equation.22.21 Z ′′ + ρ −1 Z ′ +(ω/c) 2 Z =0,withZ(a) =0<strong>and</strong>Z ′ (0) = 0; this is an SL equationwith p = ρ, q = 0 <strong>and</strong> weight function ρ/c 2 .Estimateofω 2 =[c 2 ν/(2a 2 )][0.5 −2(ν +2) −1 +(2ν +2) −1 ] −1 , which minimises to c 2 (2 + √ 2) 2 /(2a 2 )=5.83c 2 /a 2 whenν = √ 2.22.23 Note that the original equation is not self-adjoint; it needs an integrating factorof e x .Λ(y) =[ ∫ 2(1 + 0 x)ex y ′2 dx]/[ ∫ 20 ex y 2 dx; λ 0 ≤ 3/8. Since y ′ (2) must equal 0,γ =(π/2)(n + 1 ) <strong>for</strong> some integer n.222.25 E 1 ≤ (ω/2)(8n 2 +12n +3)/(4n + 1), which has a minimum value 3ω/2 wheninteger n =0.22.27 (a) V = ∫ L(p − q) dx. (c)UseV =(a − −L b)L2 to eliminate b from the expression<strong>for</strong> E; now the minimisation is with respect to a alone. The values <strong>for</strong> a <strong>and</strong> bare ±V/(2L 2 ) − Vρg/(6γ).22.29 The SL equation has p =1,q =0,<strong>and</strong>ρ =1.Use u(x, z) =x(4 − x)z(1 − z) as a trial function; numerator = 1088/90, denominator= 512/450. Direct solution k 2 =17π 2 /16.802


23Integral equationsIt is not unusual in the analysis of a physical system to encounter an equationin which an unknown but required function y(x), say, appears under an integralsign. Such an equation is called an integral equation, <strong>and</strong> in this chapter we discussseveral methods <strong>for</strong> solving the more straight<strong>for</strong>ward examples of such equations.Be<strong>for</strong>e embarking on our discussion of methods <strong>for</strong> solving various integralequations, we begin with a warning that many of the integral equations met inpractice cannot be solved by the elementary methods presented here but mustinstead be solved numerically, usually on a computer. Nevertheless, the regularoccurrence of several simple types of integral equation that may be solvedanalytically is sufficient reason to explore these equations more fully.We shall begin this chapter by discussing how a differential equation can betrans<strong>for</strong>med into an integral equation <strong>and</strong> by considering the most commontypes of linear integral equation. After introducing the operator notation <strong>and</strong>considering the existence of solutions <strong>for</strong> various types of equation, we go onto discuss elementary methods of obtaining closed-<strong>for</strong>m solutions of simpleintegral equations. We then consider the solution of integral equations in terms ofinfinite series <strong>and</strong> conclude by discussing the properties of integral equations withHermitian kernels, i.e. those in which the integr<strong>and</strong>s have particular symmetryproperties.23.1 Obtaining an integral equation from a differential equationIntegral equations occur in many situations, partly because we may always rewritea differential equation as an integral equation. It is sometimes advantageous tomake this trans<strong>for</strong>mation, since questions concerning the existence of a solutionare more easily answered <strong>for</strong> integral equations (see section 23.3), <strong>and</strong>,furthermore, an integral equation can incorporate automatically any boundaryconditions on the solution.803


INTEGRAL EQUATIONSWe shall illustrate the principles involved by considering the differential equationy ′′ (x) =f(x, y), (23.1)where f(x, y) can be any function of x <strong>and</strong> y but not of y ′ (x). Equation (23.1)thus represents a large class of linear <strong>and</strong> non-linear second-order differentialequations.We can convert (23.1) into the corresponding integral equation by first integratingwith respect to x to obtainy ′ (x) =Integrating once more, we findy(x) =∫ x0du∫ x0∫ u0f(z,y(z)) dz + c 1 .f(z,y(z)) dz + c 1 x + c 2 .Provided we do not change the region in the uz-plane over which the doubleintegral is taken, we can reverse the order of the two integrations. Changing theintegration limits appropriately, we findy(x) ==∫ x0∫ x0f(z,y(z)) dz∫ xzdu + c 1 x + c 2 (23.2)(x − z)f(z,y(z)) dz + c 1 x + c 2 ; (23.3)this is a non-linear (<strong>for</strong> general f(x, y)) Volterra integral equation.It is straight<strong>for</strong>ward to incorporate any boundary conditions on the solutiony(x) by fixing the constants c 1 <strong>and</strong> c 2 in (23.3). For example, we might have theone-point boundary condition y(0) = a <strong>and</strong> y ′ (0) = b, <strong>for</strong> which it is clear thatwe must set c 1 = b <strong>and</strong> c 2 = a.23.2 Types of integral equationFrom (23.3), we can see that even a relatively simple differential equation suchas (23.1) can lead to a corresponding integral equation that is non-linear. In thischapter, however, we will restrict our attention to linear integral equations, whichhave the general <strong>for</strong>mg(x)y(x) =f(x)+λ∫ baK(x, z)y(z) dz. (23.4)In (23.4), y(x) is the unknown function, while the functions f(x), g(x) <strong>and</strong>K(x, z)are assumed known. K(x, z) is called the kernel of the integral equation. Theintegration limits a <strong>and</strong> b are also assumed known, <strong>and</strong> may be constants orfunctions of x, <strong>and</strong>λ is a known constant or parameter.804


23.3 OPERATOR NOTATION AND THE EXISTENCE OF SOLUTIONSIn fact, we shall be concerned with various special cases of (23.4), which areknown by particular names. Firstly, if g(x) = 0 then the unknown function y(x)appears only under the integral sign, <strong>and</strong> (23.4) is called a linear integral equationof the first kind. Alternatively, if g(x) = 1, so that y(x) appears twice, once insidethe integral <strong>and</strong> once outside, then (23.4) is called a linear integral equation ofthe second kind. In either case, if f(x) = 0 the equation is called homogeneous,otherwise inhomogeneous.We can distinguish further between different types of integral equation by the<strong>for</strong>m of the integration limits a <strong>and</strong> b. If these limits are fixed constants then theequation is called a Fredholm equation. If, however, the upper limit b = x (i.e. itis variable) then the equation is called a Volterra equation; such an equation isanalogous to one with fixed limits but <strong>for</strong> which the kernel K(x, z) =0<strong>for</strong>z>x.Finally, we note that any equation <strong>for</strong> which either (or both) of the integrationlimits is infinite, or <strong>for</strong> which K(x, z) becomes infinite in the range of integration,is called a singular integral equation.23.3 Operator notation <strong>and</strong> the existence of solutionsThere is a close correspondence between linear integral equations <strong>and</strong> the matrixequations discussed in chapter 8. However, the <strong>for</strong>mer involve linear, integral relationsbetween functions in an infinite-dimensional function space (see chapter 17),whereas the latter specify linear relations among vectors in a finite-dimensionalvector space.Since we are restricting our attention to linear integral equations, it will beconvenient to introduce the linear integral operator K, whose action on anarbitrary function y is given byKy =∫ baK(x, z)y(z) dz. (23.5)This is analogous to the introduction in chapters 16 <strong>and</strong> 17 of the notation L todescribe a linear differential operator. Furthermore, we may define the Hermitianconjugate K † byK † y =∫ baK ∗ (z,x)y(z) dz,where the asterisk denotes complex conjugation <strong>and</strong> we have reversed the orderof the arguments in the kernel.It is clear from (23.5) that K is indeed linear. Moreover, since K operates onthe infinite-dimensional space of (reasonable) functions, we may make an obviousanalogy with matrix equations <strong>and</strong> consider the action of K on a function f asthat of a matrix on a column vector (both of infinite dimension).When written in operator <strong>for</strong>m, the integral equations discussed in the previoussection resemble equations familiar from linear algebra. For example, the805


INTEGRAL EQUATIONSinhomogeneous Fredholm equation of the first kind may be written as0=f + λKy,which has the unique solution y = −K −1 f/λ, provided that f ≠ 0 <strong>and</strong> the inverseoperator K −1 exists.Similarly, we may write the corresponding Fredholm equation of the secondkind asy = f + λKy. (23.6)In the homogeneous case, where f = 0, this reduces to y = λKy, which isreminiscent of an eigenvalue problem in linear algebra (except that λ appears onthe other side of the equation) <strong>and</strong>, similarly, only has solutions <strong>for</strong> at most acountably infinite set of eigenvalues λ i . The corresponding solutions y i are calledthe eigenfunctions.In the inhomogeneous case (f ≠ 0), the solution to (23.6) can be writtensymbolically asy =(1− λK) −1 f,again provided that the inverse operator exists. It may be shown that, in general,(23.6) does possess a unique solution if λ ≠ λ i ,i.e.whenλ does not equal one ofthe eigenvalues of the corresponding homogeneous equation.When λ does equal one of these eigenvalues, (23.6) may have either manysolutions or no solution at all, depending on the <strong>for</strong>m of f. If the function f isorthogonal to every eigenfunction of the equationthat belongs to the eigenvalue λ ∗ ,i.e.〈g|f〉 =g = λ ∗ K † g (23.7)∫ bag ∗ (x)f(x) dx =0<strong>for</strong> every function g obeying (23.7), then it can be shown that (23.6) has manysolutions. Otherwise the equation has no solution. These statements are discussedfurther in section 23.7, <strong>for</strong> the special case of integral equations with Hermitiankernels, i.e. those <strong>for</strong> which K = K † .23.4 Closed-<strong>for</strong>m solutionsIn certain very special cases, it may be possible to obtain a closed-<strong>for</strong>m solutionof an integral equation. The reader should realise, however, when faced with anintegral equation, that in general it will not be soluble by the simple methodspresented in this section but must instead be solved using (numerical) iterativemethods, such as those outlined in section 23.5.806


23.4 CLOSED-FORM SOLUTIONS23.4.1 Separable kernelsThe most straight<strong>for</strong>ward integral equations to solve are Fredholm equationswith separable (or degenerate) kernels. A kernel is separable if it has the <strong>for</strong>mn∑K(x, z) = φ i (x)ψ i (z), (23.8)i=1where φ i (x) areψ i (z) are respectively functions of x only <strong>and</strong> of z only <strong>and</strong> thenumber of terms in the sum, n, is finite.Let us consider the solution of the (inhomogeneous) Fredholm equation of thesecond kind,y(x) =f(x)+λ∫ baK(x, z)y(z) dz, (23.9)which has a separable kernel of the <strong>for</strong>m (23.8). Writing the kernel in its separated<strong>for</strong>m, the functions φ i (x) may be taken outside the integral over z to obtainn∑∫ by(x) =f(x)+λ φ i (x) ψ i (z)y(z) dz.i=1Since the integration limits a <strong>and</strong> b are constant <strong>for</strong> a Fredholm equation, theintegral over z in each term of the sum is just a constant. Denoting these constantsbyc i =the solution to (23.9) is found to be∫ bay(x) =f(x)+λaψ i (z)y(z) dz, (23.10)n∑c i φ i (x), (23.11)where the constants c i can be evalutated by substituting (23.11) into (23.10).◮Solve the integral equationy(x) =x + λ∫ 10i=1(xz + z 2 )y(z) dz. (23.12)The kernel <strong>for</strong> this equation is K(x, z) =xz + z 2 , which is clearly separable, <strong>and</strong> using thenotation in (23.8) we have φ 1 (x) =x, φ 2 (x) =1,ψ 1 (z) =z <strong>and</strong> ψ 2 (z) =z 2 . From (23.11)the solution to (23.12) has the <strong>for</strong>my(x) =x + λ(c 1 x + c 2 ),where the constants c 1 <strong>and</strong> c 2 are given by (23.10) asc 1 =c 2 =∫ 10∫ 10z[z + λ(c 1 z + c 2 )] dz = 1 3 + 1 3 λc 1 + 1 2 λc 2,z 2 [z + λ(c 1 z + c 2 )] dz = 1 4 + 1 4 λc 1 + 1 3 λc 2.807


INTEGRAL EQUATIONSThese two simultaneous linear equations may be straight<strong>for</strong>wardly solved <strong>for</strong> c 1 <strong>and</strong> c 2 togive24 + λ18c 1 =<strong>and</strong> c72 − 48λ − λ 2 2 =72 − 48λ − λ , 2so that the solution to (23.12) isy(x) =(72 − 24λ)x +18λ72 − 48λ − λ 2 . ◭In the above example, we see that (23.12) has a (finite) unique solution providedthat λ is not equal to either root of the quadratic in the denominator of y(x).The roots of this quadratic are in fact the eigenvalues of the correspondinghomogeneous equation, as mentioned in the previous section. In general, if theseparable kernel contains n terms, as in (23.8), there will be n such eigenvalues,although they may not all be different.Kernels consisting of trigonometric (or hyperbolic) functions of sums or differencesof x <strong>and</strong> z are also often separable.◮Find the eigenvalues <strong>and</strong> corresponding eigenfunctions of the homogeneous Fredholmequationy(x) =λ∫ π0sin(x + z) y(z) dz. (23.13)The kernel of this integral equation can be written in separated <strong>for</strong>m asK(x, z) =sin(x + z) =sinx cos z +cosx sin z,so, comparing with (23.8), we have φ 1 (x) = sinx, φ 2 (x) = cosx, ψ 1 (z) = cosz <strong>and</strong>ψ 2 (z) =sinz.Thus, from (23.11), the solution to (23.13) has the <strong>for</strong>mwhere the constants c 1 <strong>and</strong> c 2 are given byc 1 = λc 2 = λ∫ π0∫ π0y(x) =λ(c 1 sin x + c 2 cos x),cos z (c 1 sin z + c 2 cos z) dz = λπ 2 c 2, (23.14)sin z (c 1 sin z + c 2 cos z) dz = λπ 2 c 1. (23.15)Combining these two equations we find c 1 =(λπ/2) 2 c 1 , <strong>and</strong>, assuming that c 1 ≠0,thisgives λ = ±2/π, the two eigenvalues of the integral equation (23.13).By substituting each of the eigenvalues back into (23.14) <strong>and</strong> (23.15), we find thatthe eigenfunctions corresponding to the eigenvalues λ 1 =2/π <strong>and</strong> λ 2 = −2/π are givenrespectively byy 1 (x) =A(sin x +cosx) <strong>and</strong> y 2 (x) =B(sin x − cos x), (23.16)where A <strong>and</strong> B are arbitrary constants. ◭808


23.4 CLOSED-FORM SOLUTIONS23.4.2 Integral trans<strong>for</strong>m methodsIf the kernel of an integral equation can be written as a function of the differencex − z of its two arguments, then it is called a displacement kernel. An integralequation having such a kernel, <strong>and</strong> which also has the integration limits −∞ to∞, may be solved by the use of Fourier trans<strong>for</strong>ms (chapter 13).If we consider the following integral equation with a displacement kernel,y(x) =f(x)+λ∫ ∞−∞K(x − z)y(z) dz, (23.17)the integral over z clearly takes the <strong>for</strong>m of a convolution (see chapter 13).There<strong>for</strong>e, Fourier-trans<strong>for</strong>ming (23.17) <strong>and</strong> using the convolution theorem, weobtainỹ(k) =˜f(k)+ √ 2πλ ˜K(k)ỹ(k),which may be rearranged to give˜f(k)ỹ(k) =1 − √ 2πλ ˜K(k) . (23.18)Taking the inverse Fourier trans<strong>for</strong>m, the solution to (23.17) is given byy(x) = √ 1 ∫ ∞ ˜f(k) exp(ikx)2π −∞ 1 − √ 2πλ ˜K(k) dk.If we can per<strong>for</strong>m this inverse Fourier trans<strong>for</strong>mation then the solution can befound explicitly; otherwise it must be left in the <strong>for</strong>m of an integral.◮Find the Fourier trans<strong>for</strong>m of the function{1 if |x| ≤a,g(x) =0 if |x| >a.Hence find an explicit expression <strong>for</strong> the solution of the integral equation∫ ∞sin(x − z)y(x) =f(x)+λy(z) dz. (23.19)−∞ x − zFind the solution <strong>for</strong> the special case f(x) =(sinx)/x.The Fourier trans<strong>for</strong>m of g(x) is given directly by˜g(k) = √ 1 ∫ a[ ] a√1 exp(−ikx) 2 sin kaexp(−ikx) dx = √ = .2π −a2π (−ik)−aπ k(23.20)The kernel of the integral equation (23.19) is K(x − z) = [sin(x − z)]/(x − z). Using(23.20), it is straight<strong>for</strong>ward to show that the Fourier trans<strong>for</strong>m of the kernel is˜K(k) ={√π/2 if |k| ≤1,0 if |k| > 1.(23.21)809


INTEGRAL EQUATIONSThus, using (23.18), we find the Fourier trans<strong>for</strong>m of the solution to beỹ(k) ={˜f(k)/(1 − πλ) if |k| ≤1,˜f(k) if |k| > 1.(23.22)Inverse Fourier-trans<strong>for</strong>ming, <strong>and</strong> writing the result in a slightly more convenient <strong>for</strong>m,the solution to (23.19) is given by( ) ∫ 1 1 1y(x) =f(x)+1 − πλ − 1 √ ˜f(k)exp(ikx) dk2π −1= f(x)+ πλ ∫1 1√ ˜f(k)exp(ikx) dk. (23.23)1 − πλ 2π −1It is clear from (23.22) that when λ = 1/π, which is the only eigenvalue of thecorresponding homogeneous equation to (23.19), the solution becomes infinite, as wewould expect.For the special case f(x) =(sinx)/x, the Fourier trans<strong>for</strong>m ˜f(k) is identical to that in(23.21), <strong>and</strong> the solution (23.23) becomesy(x) = sin x ( ) ∫ πλ 1 1√ πx + √ exp(ikx) dk1 − πλ 2π −1 2= sin x ( [ ] k=1πλ 1 exp(ikx)x + 1 − πλ)2 ixk=−1= sin x ( ) ( ) πλ sin x 1 sin xx + 1 − πλ x = 1 − πλ x. ◭If, instead, the integral equation (23.17) had integration limits 0 <strong>and</strong> x (somaking it a Volterra equation) then its solution could be found, in a similar way,by using the convolution theorem <strong>for</strong> Laplace trans<strong>for</strong>ms (see chapter 13). Wewould find¯f(s)ȳ(s) =1 − λ ¯K(s) ,where s is the Laplace trans<strong>for</strong>m variable. Often one may use the dictionary ofLaplace trans<strong>for</strong>ms given in table 13.1 to invert this equation <strong>and</strong> find the solutiony(x). In general, however, the evaluation of inverse Laplace trans<strong>for</strong>m integralsis difficult, since (in principle) it requires a contour integration; see chapter 24.As a final example of the use of Fourier trans<strong>for</strong>ms in solving integral equations,we mention equations that have integration limits −∞ <strong>and</strong> ∞ <strong>and</strong> a kernel of the<strong>for</strong>mK(x, z) =exp(−ixz).Consider, <strong>for</strong> example, the inhomogeneous Fredholm equationy(x) =f(x)+λ∫ ∞−∞exp(−ixz) y(z) dz. (23.24)The integral over z is clearly just (a multiple of) the Fourier trans<strong>for</strong>m of y(z),810


23.4 CLOSED-FORM SOLUTIONSso we can writey(x) =f(x)+ √ 2πλỹ(x). (23.25)If we now take the Fourier trans<strong>for</strong>m of (23.25) but continue to denote theindependent variable by x (i.e. rather than k, <strong>for</strong> example), we obtainỹ(x) =˜f(x)+ √ 2πλy(−x). (23.26)Substituting (23.26) into (23.25) we findy(x) =f(x)+ √ √ ]2πλ[˜f(x)+ 2πλy(−x) ,but on making the change x →−x <strong>and</strong> substituting back in <strong>for</strong> y(−x), this givesy(x) =f(x)+ √ [ 2πλ˜f(x)+2πλ 2 f(−x)+ √ ]2πλ˜f(−x)+2πλ 2 y(x) .Thus the solution to (23.24) is given byy(x) =1[1 − (2π) 2 λ 4 f(x)+(2π) 1/2 λ˜f(x)+2πλ 2 f(−x)+(2π) 3/2 λ 3˜f(−x) ].(23.27)Clearly, (23.24) possesses a unique solution provided λ ≠ ±1/ √ 2π or ±i/ √ 2π;these are easily shown to be the eigenvalues of the corresponding homogeneousequation (<strong>for</strong> which f(x) ≡ 0).◮Solve the integral equation) ∫ ∞y(x) =exp(− x2+ λ exp(−ixz) y(z) dz, (23.28)2−∞where λ is a real constant. Show that the solution is unique unless λ has one of two particularvalues. Does a solution exist <strong>for</strong> either of these two values of λ?Following the argument given above, the solution to (23.28) is given by (23.27) withf(x) =exp(−x 2 /2). In order to write the solution explicitly, however, we must calculatethe Fourier trans<strong>for</strong>m of f(x). Using equation (13.7), we find ˜f(k) =exp(−k 2 /2), fromwhich we note that f(x) has the special property that its functional <strong>for</strong>m is identical tothat of its Fourier trans<strong>for</strong>m. Thus, the solution to (23.28) is given byy(x) =11 − (2π) 2 λ 4 [1+(2π) 1/2 λ +2πλ 2 +(2π) 3/2 λ 3] exp(− x22).(23.29)Since λ is restricted to be real, the solution to (23.28) will be unique unless λ = ±1/ √ 2π,at which points (23.29) becomes infinite. In order to find whether solutions exist <strong>for</strong> eitherof these values of λ we must return to equations (23.25) <strong>and</strong> (23.26).Let us first consider the case λ =+1/ √ 2π. Putting this value into (23.25) <strong>and</strong> (23.26),we obtainy(x) =f(x)+ỹ(x), (23.30)ỹ(x) =˜f(x)+y(−x). (23.31)811


INTEGRAL EQUATIONSSubstituting (23.31) into (23.30) we findy(x) =f(x)+˜f(x)+y(−x),but on changing x to −x <strong>and</strong> substituting back in <strong>for</strong> y(−x), this givesy(x) =f(x)+˜f(x)+f(−x)+˜f(−x)+y(x).Thus, in order <strong>for</strong> a solution to exist, we require that the function f(x) obeysf(x)+˜f(x)+f(−x)+˜f(−x) =0.This is satisfied if f(x) =−˜f(x), i.e. if the functional <strong>for</strong>m of f(x) is minus the <strong>for</strong>m of itsFourier trans<strong>for</strong>m. We may repeat this analysis <strong>for</strong> the case λ = −1/ √ 2π, <strong>and</strong>, in a similarway, we find that this time we require f(x) =˜f(x).In our case f(x) = exp(−x 2 /2), <strong>for</strong> which, as we mentioned above, f(x) = ˜f(x).There<strong>for</strong>e, (23.28) possesses no solution when λ =+1/ √ 2π but has many solutions whenλ = −1/ √ 2π. ◭A similar approach to the above may be taken to solve equations with kernelsof the <strong>for</strong>m K(x, y) =cosxy or sin xy, either by considering the integral over y ineach case as the real or imaginary part of the corresponding Fourier trans<strong>for</strong>mor by using Fourier cosine or sine trans<strong>for</strong>ms directly.23.4.3 DifferentiationA closed-<strong>for</strong>m solution to a Volterra equation may sometimes be obtained bydifferentiating the equation to obtain the corresponding differential equation,which may be easier to solve.◮Solve the integral equationy(x) =x −∫ x0xz 2 y(z) dz. (23.32)Dividing through by x, weobtain∫y(x)xx=1− z 2 y(z) dz,0which may be differentiated with respect to x to give[ ][ ]d y(x)y(x)= −x 2 y(x) =−x 3 .dx xxThis equation may be integrated straight<strong>for</strong>wardly, <strong>and</strong> we find[ ] y(x)ln = − x4x 4 + c,where c is a constant of integration. Thus the solution to (23.32) has the <strong>for</strong>m)y(x) =Ax exp(− x4, (23.33)4where A is an arbitrary constant.Since the original integral equation (23.32) contains no arbitrary constants, neithershould its solution. We may calculate the value of the constant, A, by substituting thesolution (23.33) back into (23.32), from which we find A =1.◭812


23.5 NEUMANN SERIES23.5 Neumann seriesAs mentioned above, most integral equations met in practice will not be of thesimple <strong>for</strong>ms discussed in the last section <strong>and</strong> so, in general, it is not possible tofind closed-<strong>for</strong>m solutions. In such cases, we might try to obtain a solution in the<strong>for</strong>m of an infinite series, as we did <strong>for</strong> differential equations (see chapter 16).Let us consider the equationy(x) =f(x)+λ∫ baK(x, z)y(z) dz, (23.34)where either both integration limits are constants (<strong>for</strong> a Fredholm equation) orthe upper limit is variable (<strong>for</strong> a Volterra equation). Clearly, if λ were small thena crude (but reasonable) approximation to the solution would bey(x) ≈ y 0 (x) =f(x),where y 0 (x) st<strong>and</strong>s <strong>for</strong> our ‘zeroth-order’ approximation to the solution (<strong>and</strong> isnot to be confused with an eigenfunction).Substituting this crude guess under the integral sign in the original equation,we obtain what should be a better approximation:y 1 (x) =f(x)+λ∫ baK(x, z)y 0 (z) dz = f(x)+λ∫ baK(x, z)f(z) dz,which is first order in λ. Repeating the procedure once more results in thesecond-order approximationy 2 (x) =f(x)+λ= f(x)+λ∫ ba∫ baK(x, z)y 1 (z) dz∫ b ∫ bK(x, z 1 )f(z 1 ) dz 1 + λ 2 dz 1 K(x, z 1 )K(z 1 ,z 2 )f(z 2 ) dz 2 .It is clear that we may continue this process to obtain progressively higher-orderapproximations to the solution. Introducing the functionsK 1 (x, z) =K(x, z),K 2 (x, z) =K 3 (x, z) =∫ ba∫ bK(x, z 1 )K(z 1 ,z) dz 1 ,<strong>and</strong> so on, which obey the recurrence relationaK n (x, z) =a∫ bdz 1 K(x, z 1 )K(z 1 ,z 2 )K(z 2 ,z) dz 2 ,a∫ baK(x, z 1 )K n−1 (z 1 ,z) dz 1 ,813a


INTEGRAL EQUATIONSwe may write the nth-order approximation asn∑∫ by n (x) =f(x)+ λ m K m (x, z)f(z) dz. (23.35)m=1The solution to the original integral equation is then given by y(x) =lim n→∞ y n (x), provided the infinite series converges. Using (23.35), this solutionmay be written asy(x) =f(x)+λ∫ bwhere the resolvent kernel R(x, z; λ) is given byR(x, z; λ) =aaR(x, z; λ)f(z) dz, (23.36)∞∑λ m K m+1 (x, z). (23.37)m=0Clearly, the resolvent kernel, <strong>and</strong> hence the series solution, will convergeprovided λ is sufficiently small. In fact, it may be shown that the series convergesin some domain of |λ| provided the original kernel K(x, z) is bounded in such away that∫ b ∫ b|λ| 2 dx |K(x, z)| 2 dz < 1. (23.38)aa◮Use the Neumann series method to solve the integral equationy(x) =x + λ∫ 10xzy(z) dz. (23.39)Following the method outlined above, we begin with the crude approximation y(x) ≈y 0 (x) =x. Substituting this under the integral sign in (23.39), we obtain the next approximationy 1 (x) =x + λ∫ 10xzy 0 (z) dz = x + λRepeating the procedure once more, we obtainy 2 (x) =x + λ= x + λ∫ 10∫ 10xzy 1 (z) dz∫ 1(xz z + λz )dz = x +30xz 2 dz = x + λx 3 ,( ) λ3 + λ2x.9For this simple example, it is easy to see that by continuing this process the solution to(23.39) is obtained as[ ( ) 2 ( ) 3λ λ λy(x) =x +3 + + + ···]x.3 3Clearly the expression in brackets is an infinite geometric series with first term λ/3 <strong>and</strong>814


23.6 FREDHOLM THEORYcommon ratio λ/3. Thus, provided |λ| < 3, this infinite series converges to the valueλ/(3 − λ), <strong>and</strong> the solution to (23.39) isy(x) =x +λx3 − λ = 3x3 − λ . (23.40)Finally, we note that the requirement that |λ| < 3 may also be derived very easily fromthe condition (23.38). ◭23.6 Fredholm theoryIn the previous section, we found that a solution to the integral equation (23.34)can be obtained as a Neumann series of the <strong>for</strong>m (23.36), where the resolventkernel R(x, z; λ) is written as an infinite power series in λ. This solution is validprovided the infinite series converges.A related, but more elegant, approach to the solution of integral equationsusing infinite series was found by Fredholm. We will not reproduce Fredholm’sanalysis here, but merely state the results we need. Essentially, Fredholm theoryprovides a <strong>for</strong>mula <strong>for</strong> the resolvent kernel R(x, z; λ) in (23.36) in terms of theratio of two infinite series:D(x, z; λ)R(x, z; λ) = . (23.41)d(λ)The numerator <strong>and</strong> denominator in (23.41) are given byD(x, z; λ) =d(λ) =∞∑n=0∞∑n=0(−1) nD n (x, z)λ n , (23.42)n!(−1) nd n λ n , (23.43)n!where the functions D n (x, z) <strong>and</strong> the constants d n are found from recurrencerelations as follows. We start withD 0 (x, z) =K(x, z) <strong>and</strong> d 0 =1, (23.44)where K(x, z) is the kernel of the original integral equation (23.34). The higherordercoefficients of λ in (23.43) <strong>and</strong> (23.42) are then obtained from the tworecurrence relationsd n =∫ bD n (x, z) =K(x, z)d n − naD n−1 (x, x) dx, (23.45)∫ baK(x, z 1 )D n−1 (z 1 ,z) dz 1 . (23.46)Although the <strong>for</strong>mulae <strong>for</strong> the resolvent kernel appear complicated, they areoften simple to apply. Moreover, <strong>for</strong> the Fredholm solution the power series(23.42) <strong>and</strong> (23.43) are both guaranteed to converge <strong>for</strong> all values of λ, unlike815


INTEGRAL EQUATIONSNeumann series, which converge only if the condition (23.38) is satisfied. Thus theFredholm method leads to a unique, non-singular solution, provided that d(λ) ≠0.In fact, as we might suspect, the solutions of d(λ) = 0 give the eigenvalues of thehomogeneous equation corresponding to (23.34), i.e. with f(x) ≡ 0.◮Use Fredholm theory to solve the integral equation (23.39).Using (23.36) <strong>and</strong> (23.41), the solution to (23.39) can be written in the <strong>for</strong>m∫ 1D(x, z; λ)y(x) =x + λ R(x, z; λ)zdz= x + λzdz. (23.47)00 d(λ)In order to find the <strong>for</strong>m of the resolvent kernel R(x, z; λ), we begin by settingD 0 (x, z) =K(x, z) =xz <strong>and</strong> d 0 =1<strong>and</strong> use the recurrence relations (23.45) <strong>and</strong> (23.46) to obtaind 1 =∫ 10D 0 (x, x) dx =∫ 10∫ 1x 2 dx = 1 3 ,D 1 (x, z) = xz ∫ 13 − xz1zdz 2 1 = xz [ z303 − xz 1=0.3Applying the recurrence relations again we find that d n =0<strong>and</strong>D n (x, z) =0<strong>for</strong>n>1.Thus, from (23.42) <strong>and</strong> (23.43), the numerator <strong>and</strong> denominator of the resolvent respectivelyare given byD(x, z; λ) =xz <strong>and</strong> d(λ) =1− λ 3 .Substituting these expressions into (23.47), we find that the solution to (23.39) is givenby∫ 1xz 2y(x) =x + λ1 − λ/3 dz[= x + λ0x1 − λ/3z 33] 10] 10= x + λx3 − λ = 3x3 − λ ,which, as expected, is the same as the solution (23.40) found by constructing a Neumannseries. ◭23.7 Schmidt–Hilbert theoryThe Schmidt–Hilbert (SH) theory of integral equations may be considered asanalogous to the Sturm–Liouville (SL) theory of differential equations, discussedin chapter 17, <strong>and</strong> is concerned with the properties of integral equations withHermitian kernels. An Hermitian kernel enjoys the propertyK(x, z) =K ∗ (z,x), (23.48)<strong>and</strong> it is clear that a special case of (23.48) occurs <strong>for</strong> a real kernel that is alsosymmetric with respect to its two arguments.816


23.7 SCHMIDT–HILBERT THEORYLet us begin by considering the homogeneous integral equationy = λKy,where the integral operator K has an Hermitian kernel. As discussed in section23.3, in general this equation will have solutions only <strong>for</strong> λ = λ i ,wheretheλ iare the eigenvalues of the integral equation, the corresponding solutions y i beingthe eigenfunctions of the equation.By following similar arguments to those presented in chapter 17 <strong>for</strong> SL theory,it may be shown that the eigenvalues λ i of an Hermitian kernel are real <strong>and</strong>that the corresponding eigenfunctions y i belonging to different eigenvalues areorthogonal <strong>and</strong> <strong>for</strong>m a complete set. If the eigenfunctions are suitably normalised,we have〈y i |y j 〉 =∫ bay ∗ i (x)y j(x) dx = δ ij . (23.49)If an eigenvalue is degenerate then the eigenfunctions corresponding to thateigenvalue can be made orthogonal by the Gram–Schmidt procedure, in a similarway to that discussed in chapter 17 in the context of SL theory.Like SL theory, SH theory does not provide a method of obtaining the eigenvalues<strong>and</strong> eigenfunctions of any particular homogeneous integral equation withan Hermitian kernel; <strong>for</strong> this we have to turn to the methods discussed in theprevious sections of this chapter. Rather, SH theory is concerned with the generalproperties of the solutions to such equations. Where SH theory becomesapplicable, however, is in the solution of inhomogeneous integral equations withHermitian kernels <strong>for</strong> which the eigenvalues <strong>and</strong> eigenfunctions of the correspondinghomogeneous equation are already known.Let us consider the inhomogeneous equationy = f + λKy, (23.50)where K = K † <strong>and</strong> <strong>for</strong> which we know the eigenvalues λ i <strong>and</strong> normalisedeigenfunctions y i of the corresponding homogeneous problem. The function fmay or may not be expressible solely in terms of the eigenfunctions y i ,<strong>and</strong>toaccommodate this situation we write the unknown solution y as y = f + ∑ i a iy i ,where the a i are expansion coefficients to be determined.Substituting this into (23.50), we obtainf + ∑ ia i y i = f + λ ∑ ia i y iλ i+ λKf, (23.51)wherewehaveusedthefactthaty i = λ i Ky i . Forming the inner product of both817


INTEGRAL EQUATIONSsides of (23.51) with y j , we find∑a i 〈y j |y i 〉 = λ ∑ a i〈y j |y i 〉 + λ〈y j |Kf〉. (23.52)λii iSince the eigenfunctions are orthonormal <strong>and</strong> K is an Hermitian operator,we have that both 〈y j |y i 〉 = δ ij <strong>and</strong> 〈y j |Kf〉 = 〈Ky j |f〉 = λ −1j 〈y j |f〉. Thus thecoefficients a j are given bya j = λλ−1 j 〈y j |f〉1 − λλ −1 = λ〈y j|f〉jλ j − λ , (23.53)<strong>and</strong> the solution isy = f + ∑ a i y i = f + λ ∑ 〈y i |f〉λii i − λ y i. (23.54)This also shows, incidentally, that a <strong>for</strong>mal representation <strong>for</strong> the resolvent kernelisR(x, z; λ) = ∑ iy i (x)y ∗ i (z)λ i − λ . (23.55)If f can be expressed as a linear superposition of the y i ,i.e.f = ∑ i b iy i ,thenb i = 〈y i |f〉 <strong>and</strong> the solution can be written more briefly asy = ∑ 1 − λλ −1 y i . (23.56)iiWe see from (23.54) that the inhomogeneous equation (23.50) has a uniquesolution provided λ ≠ λ i ,i.e.whenλ is not equal to one of the eigenvalues ofthe corresponding homogeneous equation. However, if λ does equal one of theeigenvalues λ j then, in general, the coefficients a j become singular <strong>and</strong> no (finite)solution exists.Returning to (23.53), we notice that even if λ = λ j a non-singular solution tothe integral equation is still possible provided that the function f is orthogonalto every eigenfunction corresponding to the eigenvalue λ j ,i.e.〈y j |f〉 =∫ bab iy ∗ j (x)f(x) dx =0.The following worked example illustrates the case in which f can be expressed interms of the y i . One in which it cannot is considered in exercise 23.14.◮Use Schmidt–Hilbert theory to solve the integral equationy(x) = sin(x + α)+λ∫ π0sin(x + z)y(z) dz. (23.57)It is clear that the kernel K(x, z) =sin(x + z) is real <strong>and</strong> symmetric in x <strong>and</strong> z <strong>and</strong> is818


23.8 EXERCISESthus Hermitian. In order to solve this inhomogeneous equation using SH theory, however,we must first find the eigenvalues <strong>and</strong> eigenfunctions of the corresponding homogeneousequation.In fact, we have considered the solution of the corresponding homogeneous equation(23.13) already, in subsection 23.4.1, where we found that it has two eigenvalues λ 1 =2/π<strong>and</strong> λ 2 = −2/π, with eigenfunctions given by (23.16). The normalised eigenfunctions arey 1 (x) = √ 1 (sin x +cosx) <strong>and</strong> y 2 (x) = √ 1 (sin x − cos x) (23.58)π π<strong>and</strong> are easily shown to obey the orthonormality condition (23.49).Using (23.54), the solution to the inhomogeneous equation (23.57) has the <strong>for</strong>my(x) =a 1 y 1 (x)+a 2 y 2 (x), (23.59)where the coefficients a 1 <strong>and</strong> a 2 are given by (23.53) with f(x) =sin(x + α). There<strong>for</strong>e,using (23.58),1a 1 =1 − πλ/21a 2 =1+πλ/2∫ π0∫ π01√ π(sin z +cosz) sin(z + α) dz =1√ π(sin z − cos z) sin(z + α) dz =√ π(cos α +sinα),2 − πλ√ π(cos α − sin α).2+πλSubstituting these expressions <strong>for</strong> a 1 <strong>and</strong> a 2 into (23.59) <strong>and</strong> simplifying, we find thatthe solution to (23.57) is given byy(x) =11 − (πλ/2) 2 [sin(x + α)+(πλ/2) cos(x − α)]. ◭23.1 Solve the integral equation∫ ∞023.8 Exercisescos(xv)y(v) dv =exp(−x 2 /2)<strong>for</strong> the function y = y(x) <strong>for</strong>x>0. Note that <strong>for</strong> x


INTEGRAL EQUATIONS23.5 Solve <strong>for</strong> φ(x) the integral equationφ(x) =f(x)+λ∫ 10[( xy) n ( y) ] n+ φ(y) dy,xwhere f(x) is bounded <strong>for</strong> 0


23.8 EXERCISES(b) Obtain the eigenvalues <strong>and</strong> eigenfunctions over the interval [0, 2π] if∞∑ 1K(x, y) = cos nx cos ny.nn=123.8 By taking its Laplace trans<strong>for</strong>m, <strong>and</strong> that of x n e −ax , obtain the explicit solutionof∫ x]f(x) =e[x −x + (x − u)e u f(u) du .Verify your answer by substitution.23.9 For f(t) =exp(−t 2 /2), use the relationships of the Fourier trans<strong>for</strong>ms of f ′ (t) <strong>and</strong>tf(t) tothatoff(t) itself to find a simple differential equation satisfied by ˜f(ω),the Fourier trans<strong>for</strong>m of f(t), <strong>and</strong> hence determine ˜f(ω) to within a constant.Use this result to solve the integral equation<strong>for</strong> h(t).23.10 Show that the equation∫ ∞−∞f(x) =x −1/3 + λ0e −t(t−2x)/2 h(t) dt = e 3x2 /8∫ ∞0f(y)exp(−xy) dyhas a solution of the <strong>for</strong>m Ax α + Bx β . Determine the values of α <strong>and</strong> β, <strong>and</strong>showthat those of A <strong>and</strong> B are11 − λ 2 Γ( 1 3 )Γ( 2 3 ) <strong>and</strong>λΓ( 2 3 )1 − λ 2 Γ( 1 3 )Γ( 2 3 ) ,where Γ(z) is the gamma function.23.11 At an international ‘peace’ conference a large number of delegates are seatedaround a circular table with each delegation sitting near its allies <strong>and</strong> diametricallyopposite the delegation most bitterly opposed to it. The position of a delegate isdenoted by θ, with0≤ θ ≤ 2π. Thefuryf(θ) felt by the delegate at θ is the sumof his own natural hostility h(θ) <strong>and</strong> the influences on him of each of the otherdelegates; a delegate at position φ contributes an amount K(θ − φ)f(φ). Thusf(θ) =h(θ)+∫ 2π0K(θ − φ)f(φ) dφ.Show that if K(ψ) takesthe<strong>for</strong>mK(ψ) =k 0 + k 1 cos ψ thenf(θ) =h(θ)+p + q cos θ + r sin θ<strong>and</strong> evaluate p, q <strong>and</strong> r. A positive value <strong>for</strong> k 1 implies that delegates tend toplacate their opponents but upset their allies, whilst negative values imply thatthey calm their allies but infuriate their opponents. A walkout will occur if f(θ)exceeds a certain threshold value <strong>for</strong> some θ. Isthismorelikelytohappen<strong>for</strong>positive or <strong>for</strong> negative values of k 1 ?23.12 By considering functions of the <strong>for</strong>m h(x) = ∫ x(x − y)f(y) dy, show that the0solution f(x) oftheintegralequationf(x) =x + 1 2satisfies the equation f ′′ (x) =f(x).821∫ 10|x − y|f(y) dy


INTEGRAL EQUATIONSBy examining the special cases x =0<strong>and</strong>x = 1, show that2f(x) =(e +3)(e +1) [(e +2)ex − ee −x ].23.13 The operator M is defined byMf(x) ≡∫ ∞−∞K(x, y)f(y) dy,where K(x, y) = 1 inside the square |x|


23.9 HINTS AND ANSWERS23.9 Hints <strong>and</strong> answers23.1 Define y(−x) =y(x) <strong>and</strong> use the cosine Fourier trans<strong>for</strong>m inversion theorem;y(x) =(2/π) 1/2 exp(−x 2 /2).23.3 f ′′ (x) − f(x) =expx; α =3/4, β =1/2, γ =1/4.23.5 (a) φ(x) =f(x) − (1+2n)F n x n − (1 − 2n)F −n x −n . (b) There are no solutions <strong>for</strong>λ =[1± (1 − 4n 2 ) −1/2 ] −1 unless F ±n =0orF n /F −n = ∓[(1 − 2n)/(1 + 2n)] 1/2 .23.7 (a) a (i)n= ∫ ba h n(x)ψ (i) (x) dx; (b) use (1/ √ π)cosnx <strong>and</strong> (1/ √ π)sinnx; M is diagonal;eigenvalues λ k = k/π with eigenfunctions ψ (k) (x) =(1/ √ π)coskx.23.9 d˜f/dω = −ω˜f, leading to ˜f(ω) =Ae −ω2 /2 . Rearrange the integral as a convolution<strong>and</strong> deduce that ˜h(ω) = Be −3ω2 /2 ; h(t) = Ce −t2 /6 , where resubstitution <strong>and</strong>Gaussian normalisation show that C = √ 2/(3π).23.11 p = k 0 H/(1 − 2πk 0 ), q = k 1 H c /(1 − πk 1 )<strong>and</strong>r = k 1 H s /(1 − πk 1 ),where H = ∫ 2πh(z) dz, H0 c = ∫ 2πh(z)coszdz,<strong>and</strong>H0 s = ∫ 2πh(z)sinzdz. Positive0values of k 1 (≈ π −1 ) are most likely to cause a conference breakdown.23.13 For eigenvalue 0 : f(x) =0<strong>for</strong>|x|


24Complex variablesThroughout this book references have been made to results derived from the theoryof complex variables. This theory thus becomes an integral part of the mathematicsappropriate to physical applications. Indeed, so numerous <strong>and</strong> widespreadare these applications that the whole of the next chapter is devoted to a systematicpresentation of some of the more important ones. This current chapter developsthe general theory on which these applications are based. The difficulty with it,from the point of view of a book such as the present one, is that the underlyingbasis has a distinctly pure mathematics flavour.Thus, to adopt a comprehensive rigorous approach would involve a largeamount of groundwork in analysis, <strong>for</strong> example <strong>for</strong>mulating precise definitionsof continuity <strong>and</strong> differentiability, developing the theory of sets <strong>and</strong> making adetailed study of boundedness. Instead, we will be selective <strong>and</strong> pursue only thoseparts of the <strong>for</strong>mal theory that are needed to establish the results used in the nextchapter <strong>and</strong> elsewhere in this book.In this spirit, the proofs that have been adopted <strong>for</strong> some of the st<strong>and</strong>ardresults of complex variable theory have been chosen with an eye to simplicityrather than sophistication. This means that in some cases the imposed conditionsare more stringent than would be strictly necessary if more sophisticated proofswere used; where this happens the less restrictive results are usually stated aswell. The reader who is interested in a fuller treatment should consult one of themany excellent textbooks on this fascinating subject. §One further concession to ‘h<strong>and</strong>-waving’ has been made in the interests ofkeeping the treatment to a moderate length. In several places phrases such as ‘canbe made as small as we like’ are used, rather than a careful treatment in termsof ‘given ɛ>0, there exists a δ>0 such that’. In the authors’ experience, some§ For example, K. Knopp, Theory of Functions, Part I (New York: Dover, 1945); E. G. Phillips,Functions of a Complex Variable with Applications 7th edn (Edinburgh: Oliver <strong>and</strong> Boyd, 1951); E.C. Titchmarsh, The Theory of Functions (Ox<strong>for</strong>d: Ox<strong>for</strong>d University Press, 1952).824


24.1 FUNCTIONS OF A COMPLEX VARIABLEstudents are more at ease with the <strong>for</strong>mer type of statement, despite its lack ofprecision, whilst others, those who would contemplate only the latter, are usuallywell able to supply it <strong>for</strong> themselves.24.1 Functions of a complex variableThe quantity f(z) is said to be a function of the complex variable z if to every valueof z in a certain domain R (a region of the Arg<strong>and</strong> diagram) there correspondsone or more values of f(z). Stated like this f(z) could be any function consistingof a real <strong>and</strong> an imaginary part, each of which is, in general, itself a function of x<strong>and</strong> y. If we denote the real <strong>and</strong> imaginary parts of f(z) byu <strong>and</strong> v, respectively,thenf(z) =u(x, y)+iv(x, y).In this chapter, however, we will be primarily concerned with functions that aresingle-valued, so that to each value of z there corresponds just one value of f(z),<strong>and</strong> are differentiable in a particular sense, which we now discuss.A function f(z) that is single-valued in some domain R is differentiable at thepoint z in R if the derivative[ ]f(z +∆z) − f(z)f ′ (z) = lim(24.1)∆z→0 ∆zexists <strong>and</strong> is unique, in that its value does not depend upon the direction in theArg<strong>and</strong> diagram from which ∆z tends to zero.◮Show that the function f(z) =x 2 − y 2 + i2xy is differentiable <strong>for</strong> all values of z.Considering the definition (24.1), <strong>and</strong> taking ∆z =∆x + i∆y, we havef(z +∆z) − f(z)∆z= (x +∆x)2 − (y +∆y) 2 +2i(x +∆x)(y +∆y) − x 2 + y 2 − 2ixy∆x + i∆y= 2x∆x +(∆x)2 − 2y∆y − (∆y) 2 +2i(x∆y + y∆x +∆x∆y)∆x + i∆y=2x + i2y + (∆x)2 − (∆y) 2 +2i∆x∆y.∆x + i∆yNow, in whatever way ∆x <strong>and</strong> ∆y are allowed to tend to zero (e.g. taking ∆y =0<strong>and</strong>letting ∆x → 0 or vice versa), the last term on the RHS will tend to zero <strong>and</strong> the uniquelimit 2x + i2y will be obtained. Since z was arbitrary, f(z) withu = x 2 − y 2 <strong>and</strong> v =2xyis differentiable at all points in the (finite) complex plane. ◭We note that the above working can be considerably reduced by recognisingthat, since z = x + iy, we can write f(z) asf(z) =x 2 − y 2 +2ixy =(x + iy) 2 = z 2 .825


COMPLEX VARIABLESWe then find that[ (z +∆z)f ′ 2 − z 2 ] [ (∆z) 2 ]+2z∆z(z) = lim= lim∆z→0 ∆z∆z→0 ∆z( )= lim ∆z +2z =2z,∆z→0from which we see immediately that the limit both exists <strong>and</strong> is independent ofthe way in which ∆z → 0. Thus we have verified that f(z) =z 2 is differentiable<strong>for</strong> all (finite) z. We also note that the derivative is analogous to that found <strong>for</strong>real variables.Although the definition of a differentiable function clearly includes a wideclass of functions, the concept of differentiability is restrictive <strong>and</strong>, indeed, somefunctions are not differentiable at any point in the complex plane.◮Show that the function f(z) =2y +ix is not differentiable anywhere in the complex plane.In this case f(z) cannot be written simply in terms of z, <strong>and</strong> so we must consider thelimit (24.1) in terms of x <strong>and</strong> y explicitly. Following the same procedure as in the previousexample we findf(z +∆z) − f(z)∆z2y +2∆y + ix + i∆x − 2y − ix=∆x + i∆y2∆y + i∆x=∆x + i∆y .In this case the limit will clearly depend on the direction from which ∆z → 0. Suppose∆z → 0 along a line through z of slope m, sothat∆y = m∆x, thenlim∆z→0[ f(z +∆z) − f(z)∆z] [ ]2∆y + i∆x= lim= 2m + i∆x, ∆y→0 ∆x + i∆y 1+im .This limit is dependent on m <strong>and</strong> hence on the direction from which ∆z → 0. Since thisconclusion is independent of the value of z, <strong>and</strong> hence true <strong>for</strong> all z, f(z) =2y + ix isnowhere differentiable. ◭A function that is single-valued <strong>and</strong> differentiable at all points of a domain Ris said to be analytic (or regular) inR. A function may be analytic in a domainexcept at a finite number of points (or an infinite number if the domain isinfinite); in this case it is said to be analytic except at these points, which arecalled the singularities of f(z). In our treatment we will not consider cases inwhich an infinite number of singularities occur in a finite domain.826


24.2 THE CAUCHY–RIEMANN RELATIONS◮Show that the function f(z) =1/(1 − z) is analytic everywhere except at z =1.Since f(z) is given explicitly as a function of z, evaluation of the limit (24.1) is somewhateasier. We find[ ]f(z +∆z) − f(z)f ′ (z) = lim∆z→0 ∆z[ ( 1 1= lim∆z→0 ∆z 1 − z − ∆z − 1 )]1 − z= lim∆z→0[1(1 − z − ∆z)(1 − z)]=1(1 − z) 2 ,independently of the way in which ∆z → 0, provided z ≠ 1. Hence f(z) is analyticeverywhere except at the singularity z =1.◭24.2 The Cauchy–Riemann relationsFrom examining the previous examples, it is apparent that <strong>for</strong> a function f(z)to be differentiable <strong>and</strong> hence analytic there must be some particular connectionbetween its real <strong>and</strong> imaginary parts u <strong>and</strong> v.By considering a general function we next establish what this connection mustbe. If the limit[ ]f(z +∆z) − f(z)L = lim(24.2)∆z→0 ∆zis to exist <strong>and</strong> be unique, in the way required <strong>for</strong> differentiability, then any twospecific ways of letting ∆z → 0 must produce the same limit. In particular, movingparallel to the real axis <strong>and</strong> moving parallel to the imaginary axis must do so.This is certainly a necessary condition, although it may not be sufficient.If we let f(z) =u(x, y)+iv(x, y) <strong>and</strong>∆z =∆x + i∆y then we havef(z +∆z) =u(x +∆x, y +∆y)+iv(x +∆x, y +∆y),<strong>and</strong> the limit (24.2) is given by[ u(x +∆x, y +∆y)+iv(x +∆x, y +∆y) − u(x, y) − iv(x, y)L = lim∆x, ∆y→0∆x + i∆yIf we first suppose that ∆z is purely real, so that ∆y = 0, we obtain[ u(x +∆x, y) − u(x, y)L = lim+ i∆x→0 ∆xv(x +∆x, y) − v(x, y)∆x].]= ∂u∂x + i ∂v∂x , (24.3)provided each limit exists at the point z. Similarly, if ∆z is taken as purelyimaginary, so that ∆x = 0, we find[ ]u(x, y +∆y) − u(x, y) v(x, y +∆y) − v(x, y)L = lim+ i = 1 ∂u∆y→0 i∆yi∆yi ∂y + ∂v∂y . (24.4)827


COMPLEX VARIABLESFor f to be differentiable at the point z, expressions (24.3) <strong>and</strong> (24.4) mustbe identical. It follows from equating real <strong>and</strong> imaginary parts that necessaryconditions <strong>for</strong> this are∂u∂x = ∂v∂y<strong>and</strong>∂v∂x = − ∂u∂y . (24.5)These two equations are known as the Cauchy–Riemann relations.We can now see why <strong>for</strong> the earlier examples (i) f(z) =x 2 − y 2 + i2xy mightbe differentiable <strong>and</strong> (ii) f(z) =2y + ix could not be.(i) u = x 2 − y 2 , v =2xy:∂u ∂v=2x = <strong>and</strong>∂x ∂y(ii) u =2y, v = x:∂u ∂v=0= but∂x ∂y∂v ∂u=2y = −∂x ∂y ,∂v∂u=1≠ −2 =−∂x ∂y .It is apparent that <strong>for</strong> f(z) to be analytic something more than the existenceof the partial derivatives of u <strong>and</strong> v with respect to x <strong>and</strong> y is required; thissomething is that they satisfy the Cauchy–Riemann relations.We may enquire also as to the sufficient conditions <strong>for</strong> f(z) to be analytic inR. It can be shown § that a sufficient condition is that the four partial derivativesexist, are continuous <strong>and</strong> satisfy the Cauchy–Riemann relations. It is the additionalrequirement of continuity that makes the difference between the necessaryconditions <strong>and</strong> the sufficient conditions.◮In which domain(s) of the complex plane is f(z) =|x|−i|y| an analytic function?Writing f = u + iv it is clear that both ∂u/∂y <strong>and</strong> ∂v/∂x arezeroinallfourquadrants<strong>and</strong> hence that the second Cauchy–Riemann relation in (24.5) is satisfied everywhere.Turning to the first Cauchy–Riemann relation, in the first quadrant (x >0, y>0) wehave f(z) =x − iy so that∂u∂x =1,∂v∂y = −1,which clearly violates the first relation in (24.5). Thus f(z) is not analytic in the firstquadrant.Following a similiar argument <strong>for</strong> the other quadrants, we find∂u= −1∂xor +1 <strong>for</strong> x0, respectively,∂v= −1∂yor +1 <strong>for</strong> y>0<strong>and</strong>y


24.2 THE CAUCHY–RIEMANN RELATIONSSince x <strong>and</strong> y are related to z <strong>and</strong> its complex conjugate z ∗ byx = 1 2 (z + z∗ ) <strong>and</strong> y = 1 2i (z − z∗ ), (24.6)we may <strong>for</strong>mally regard any function f = u + iv as a function of z <strong>and</strong> z ∗ , ratherthan x <strong>and</strong> y. If we do this <strong>and</strong> examine ∂f/∂z ∗ we obtain∂f∂z ∗ = ∂f ∂x∂x ∂z ∗ + ∂f ∂y∂y ∂z ∗( ∂u=∂x ∂x)( + i ∂v ) 12= 1 ( ∂u2 ∂x − ∂v )+ i ∂y 2( ∂u+∂y + i ∂v∂y( ∂v∂x + ∂u∂y)(− 1 )2i). (24.7)Now, if f is analytic then the Cauchy–Riemann relations (24.5) must be satisfied,<strong>and</strong> these immediately give that ∂f/∂z ∗ is identically zero. Thus we conclude thatif f is analytic then f cannot be a function of z ∗ <strong>and</strong> any expression representingan analytic function of z can contain x <strong>and</strong> y only in the combination x + iy, notin the combination x − iy.We conclude this section by discussing some properties of analytic functionsthat are of great practical importance in theoretical physics. These can be obtainedsimply from the requirement that the Cauchy–Riemann relations must be satisfiedby the real <strong>and</strong> imaginary parts of an analytic function.The most important of these results can be obtained by differentiating thefirst Cauchy–Riemann relation with respect to one independent variable, <strong>and</strong> thesecond with respect to the other independent variable, to obtain the two chainsof equalities(∂ ∂u∂x ∂x(∂ ∂v∂x ∂x)= ∂∂x)= − ∂∂x( ∂v∂y)= ∂∂y( ∂u∂y)( ∂v∂x= − ∂∂y)= − ∂∂y( ) ∂u∂x( ∂u∂y= − ∂∂y),( ∂v∂yThus both u <strong>and</strong> v are separately solutions of Laplace’s equation in two dimensions,i.e.∂ 2 u∂x 2 + ∂2 u∂y 2 = 0 <strong>and</strong> ∂ 2 v∂x 2 + ∂2 v=0. (24.8)∂y2 We will make significant use of this result in the next chapter.A further useful result concerns the two families of curves u(x, y) = constant<strong>and</strong> v(x, y) = constant, where u <strong>and</strong> v are the real <strong>and</strong> imaginary parts of anyanalytic function f = u + iv. As discussed in chapter 10, the vector normal to thecurve u(x, y) = constant is given by∇u = ∂u∂x i + ∂u j, (24.9)∂y829).


COMPLEX VARIABLESwhere i <strong>and</strong> j are the unit vectors along the x- <strong>and</strong>y-axes, respectively. A similarexpression exists <strong>for</strong> ∇v, the normal to the curve v(x, y) = constant. Taking thescalar product of these two normal vectors, we obtain∇u · ∇v = ∂u ∂v∂x ∂x + ∂u ∂v∂y ∂y= − ∂u ∂u∂x ∂y + ∂u ∂u∂y ∂x =0,where in the last line we have used the Cauchy–Riemann relations to rewritethe partial derivatives of v as partial derivatives of u. Since the scalar productof the normal vectors is zero, they must be orthogonal, <strong>and</strong> the curves u(x, y) =constant <strong>and</strong> v(x, y) = constant must there<strong>for</strong>e intersect at right angles.◮Use the Cauchy–Riemann relations to show that, <strong>for</strong> any analytic function f = u + iv, therelation |∇u| = |∇v| must hold.From (24.9) we have( ) 2 ( ) 2 ∂u ∂u|∇u| 2 = ∇u · ∇u = + .∂x ∂yUsing the Cauchy–Riemann relations to write the partial derivatives of u in terms of thoseof v, weobtain( ) 2 ( ) 2 ∂v ∂v|∇u| 2 = + = |∇v| 2 ,∂y ∂xfrom which the result |∇u| = |∇v| follows immediately. ◭24.3 Power series in a complex variableThe theory of power series in a real variable was considered in chapter 4, whichalso contained a brief discussion of the natural extension of this theory to a seriessuch as∞∑f(z) = a n z n , (24.10)n=0where z is a complex variable <strong>and</strong> the a n are, in general, complex. We nowconsider complex power series in more detail.Expression (24.10) is a power series about the origin <strong>and</strong> may be used <strong>for</strong>general discussion, since a power series about any other point z 0 can be obtainedby a change of variable from z to z − z 0 .Ifz were written in its modulus <strong>and</strong>argument <strong>for</strong>m, z = r exp iθ, expression (24.10) would becomef(z) =∞∑a n r n exp(inθ). (24.11)n=0830


24.3 POWER SERIES IN A COMPLEX VARIABLEThis series is absolutely convergent if∞∑|a n |r n , (24.12)n=0which is a series of positive real terms, is convergent. Thus tests <strong>for</strong> the absoluteconvergence of real series can be used in the present context, <strong>and</strong> of these themost appropriate <strong>for</strong>m is based on the Cauchy root test. With the radius ofconvergence R defined by1R = lim |a n| 1/n , (24.13)n→∞the series (24.10) is absolutely convergent if |z| R.If |z| = R then no particular conclusion may be drawn, <strong>and</strong> this case must beconsidered separately, as discussed in subsection 4.5.1.A circle of radius R centred on the origin is called the circle of convergenceof the series ∑ a n z n .ThecasesR = 0 <strong>and</strong> R = ∞ correspond, respectively, toconvergence at the origin only <strong>and</strong> convergence everywhere. For R finite theconvergence occurs in a restricted part of the z-plane (the Arg<strong>and</strong> diagram). Fora power series about a general point z 0 , the circle of convergence is, of course,centred on that point.◮Find the parts of the z-plane <strong>for</strong> which the following series are convergent:∞∑ z n∞(i)n! , (ii) ∑∞∑n!z n z n, (iii)n .n=0n=0n=1(i) Since (n!) 1/n behaves like n as n →∞we find lim(1/n!) 1/n =0.HenceR = ∞ <strong>and</strong> theseries is convergent <strong>for</strong> all z. (ii) Correspondingly, lim(n!) 1/n = ∞. Thus R =0<strong>and</strong>theseries converges only at z = 0. (iii) As n →∞,(n) 1/n has a lower limit of 1 <strong>and</strong> hencelim(1/n) 1/n =1/1 = 1. Thus the series is absolutely convergent if the condition |z| < 1issatisfied. ◭Case (iii) in the above example provides a good illustration of the fact thaton its circle of convergence a power series may or may not converge. For thisparticular series, the circle of convergence is |z| = 1, so let us consider theconvergence of the series at two different points on this circle. Taking z =1,theseries becomes∞∑ 1n =1+1 2 + 1 3 + 1 4 + ··· ,n=1which is easily shown to diverge (by, <strong>for</strong> example, grouping terms, as discussed insubsection 4.3.2). Taking z = −1, however, the series is given by∞∑ (−1) n= −1+ 1 n2 − 1 3 + 1 4 − ··· ,n=1831


COMPLEX VARIABLESwhich is an alternating series whose terms decrease in magnitude <strong>and</strong> whichthere<strong>for</strong>e converges.The ratio test discussed in subsection 4.3.2 may also be employed to investigatethe absolute convergence of a complex power series. A series is absolutelyconvergent if|a n+1 ||z| n+1 |a n+1 ||z|limn→∞ |a n ||z| n = lim < 1 (24.14)n→∞ |a n |<strong>and</strong> hence the radius of convergence R of the series is given by1R = limn→∞|a n+1 |.|a n |For instance, in case (i) of the previous example, we have1R = lim n!n→∞ (n +1)! = lim 1n→∞ n +1 =0.Thus the series is absolutely convergent <strong>for</strong> all (finite) z, confirming the previousresult.Be<strong>for</strong>e turning to particular power series, we conclude this section by statingthe important result § that the power series ∑ ∞0 a nz n has a sum that is an analyticfunction of z inside its circle of convergence.As a corollary to the above theorem, it may further be shown that if f(z) =∑an z n then, inside the circle of convergence of the series,f ′ (z) =∞∑na n z n−1 .n=0Repeated application of this result demonstrates that any power series can bedifferentiated any number of times inside its circle of convergence.24.4 Some elementary functionsIn the example at the end of the previous section it was shown that the functionexp z defined by∞∑ z nexp z =(24.15)n!n=0is convergent <strong>for</strong> all z of finite modulus <strong>and</strong> is thus, by the discussion ofthe previous section, an analytic function over the whole z-plane. Like its§ For a proof see, <strong>for</strong> example, K. F. Riley, <strong>Mathematical</strong> <strong>Methods</strong> <strong>for</strong> the Physical Sciences (Cambridge:Cambridge University Press, 1974), p. 446. Functions that are analytic in the whole z-plane are usually called integral or entire functions.832


24.4 SOME ELEMENTARY FUNCTIONSreal-variable counterpart it is called the exponential function; also like its realcounterpart it is equal to its own derivative.The multiplication of two exponential functions results in a further exponentialfunction, in accordance with the corresponding result <strong>for</strong> real variables.◮Show that exp z 1 exp z 2 =exp(z 1 + z 2 ).From the series expansion (24.15) of exp z 1 <strong>and</strong> a similar expansion <strong>for</strong> exp z 2 ,itisclearthat the coefficient of z1 rzs 2 in the corresponding series expansion of exp z 1 exp z 2 is simply1/(r!s!).But, from (24.15) we also have∞∑ (z 1 + z 2 ) nexp(z 1 + z 2 )=.n!n=0In order to find the coefficient of z1 rzs 2 in this expansion, we clearly have to consider theterm in which n = r + s, namely(z 1 + z 2 ) r+s 1 (=r+s)C 0 z r+s1+ ···+ r+s C s z r(r + s)! (r + s)!1z2 s + ···+ r+s C r+s z r+s2 .The coefficient of z1 rzs 2 in this is given byr+s 1 (r + s)! 1C s =(r + s)! s!r! (r + s)! = 1r!s! .Thus, since the corresponding coefficients on the two sides are equal, <strong>and</strong> all the seriesinvolved are absolutely convergent <strong>for</strong> all z, we can conclude that exp z 1 exp z 2 =exp(z 1 +z 2 ). ◭As an extension of (24.15) we may also define the complex exponent of a realnumber a>0 by the equationa z = exp(z ln a), (24.16)where ln a is the natural logarithm of a. The particular case a = e <strong>and</strong> the factthat ln e = 1 enable us to write exp z interchangeably with e z .Ifz is real then thedefinition agrees with the familiar one.The result <strong>for</strong> z = iy,exp iy =cosy + i sin y, (24.17)has been met already in equation (3.23). Its immediate extension isexp z =(expx)(cos y + i sin y). (24.18)As z varies over the complex plane, the modulus of exp z takes all real positivevalues, except that of 0. However, two values of z that differ by 2πki, <strong>for</strong> anyinteger k, produce the same value of exp z, as given by (24.18), <strong>and</strong> so exp z isperiodic with period 2πi. If we denote exp z by t, then the strip −π


COMPLEX VARIABLESderived from them (e.g. tan <strong>and</strong> tanh), the identities they satisfy <strong>and</strong> theirderivative properties are also just as <strong>for</strong> real variables. In view of this we will notgive them further attention here.The inverse function of exp z is given by w, the solution ofexp w = z. (24.19)This inverse function was discussed in chapter 3, but we mention it again here<strong>for</strong> completeness. By virtue of the discussion following (24.18), w is not uniquelydefined <strong>and</strong> is indeterminate to the extent of any integer multiple of 2πi. Ifweexpress z asz = r exp iθ,where r is the (real) modulus of z <strong>and</strong> θ is its argument (−π


24.5 MULTIVALUED FUNCTIONS AND BRANCH CUTSOn the RHS let us write t as follows:t = r exp[i(θ +2kπ)],where k is an integer. We then obtain[ ]1t 1/n (θ +2kπ)=exp ln r + in n[ ]= r 1/n (θ +2kπ)exp i ,nwhere k =0, 1,...,n− 1; <strong>for</strong> other values of k we simply recover the roots already found.Thus t has n distinct nth roots. ◭24.5 Multivalued functions <strong>and</strong> branch cutsIn the definition of an analytic function, one of the conditions imposed wasthat the function is single-valued. However, as shown in the previous section, thelogarithmic function, a complex power <strong>and</strong> a complex root are all multivalued.Nevertheless, it happens that the properties of analytic functions can still beapplied to these <strong>and</strong> other multivalued functions of a complex variable providedthat suitable care is taken. This care amounts to identifying the branch points ofthe multivalued function f(z) in question. If z is varied in such a way that itspath in the Arg<strong>and</strong> diagram <strong>for</strong>ms a closed curve that encloses a branch point,then, in general, f(z) will not return to its original value.For definiteness let us consider the multivalued function f(z) =z 1/2 <strong>and</strong> expressz as z = r exp iθ. From figure 24.1(a), it is clear that, as the point z traverses anyclosed contour C that does not enclose the origin, θ will return to its originalvalue after one complete circuit. However, <strong>for</strong> any closed contour C ′ that doesenclose the origin, after one circuit θ → θ +2π (see figure 24.1(b)). Thus, <strong>for</strong> thefunction f(z) =z 1/2 , after one circuitr 1/2 exp(iθ/2) → r 1/2 exp[i(θ +2π)/2] = −r 1/2 exp(iθ/2).In other words, the value of f(z) changes around any closed loop enclosing theorigin; in this case f(z) →−f(z). Thus z = 0 is a branch point of the functionf(z) =z 1/2 .We note in this case that if any closed contour enclosing the origin is traversedtwice then f(z) =z 1/2 returns to its original value. The number of loops arounda branch point required <strong>for</strong> any given function f(z) to return to its original valuedepends on the function in question, <strong>and</strong> <strong>for</strong> some functions (e.g. Ln z, which alsohas a branch point at the origin) the original value is never recovered.In order that f(z) may be treated as single-valued, we may define a branch cutin the Arg<strong>and</strong> diagram. A branch cut is a line (or curve) in the complex plane<strong>and</strong> may be regarded as an artificial barrier that we must not cross. Branch cutsare positioned in such a way that we are prevented from making a complete835


COMPLEX VARIABLESy Cy yrθxrθxxC ′(a) (b) (c)Figure 24.1 (a) A closed contour not enclosing the origin; (b) a closedcontour enclosing the origin; (c) a possible branch cut <strong>for</strong> f(z) =z 1/2 .circuit around any one branch point, <strong>and</strong> so the function in question remainssingle-valued.For the function f(z) =z 1/2 , we may take as a branch cut any curve startingat the origin z = 0 <strong>and</strong> extending out to |z| = ∞ in any direction, since allsuch curves would equally well prevent us from making a closed loop aroundthe branch point at the origin. It is usual, however, to take the cut along eitherthe real or the imaginary axis. For example, in figure 24.1(c), we take the cut asthe positive real axis. By agreeing not to cross this cut, we restrict θ to lie in therange 0 ≤ θ


24.6 SINGULARITIES AND ZEROS OF COMPLEX FUNCTIONSyzyyir 1θ 1r 2iixxx−iθ 2−i−i(a) (b) (c)Figure 24.2 (a) Coordinates used in the analysis of the branch points off(z) =(z 2 +1) 1/2 ; (b) one possible arrangement of branch cuts; (c) anotherpossible branch cut, which is finite.(iii) z = −i but not z = i, thenθ 1 → θ 1 , θ 2 → θ 2 +2π <strong>and</strong> so f(z) →−f(z);(iv) both branch points, then θ 1 → θ 1 +2π, θ 2 → θ 2 +2π <strong>and</strong> so f(z) → f(z).Thus, as expected, f(z) changes value around loops containing either z = i or z = −i (butnot both). We must there<strong>for</strong>e choose branch cuts that prevent us from making a completeloop around either branch point; one suitable choice is shown in figure 24.2(b).For this f(z), however, we have noted that after traversing a loop containing both branchpoints the function returns to its original value. Thus we may choose an alternative, finite,branch cut that allows this possibility but still prevents us from making a complete looparound just one of the points. A suitable cut is shown in figure 24.2(c). ◭24.6 Singularities <strong>and</strong> zeros of complex functionsA singular point of a complex function f(z) is any point in the Arg<strong>and</strong> diagramat which f(z) fails to be analytic. We have already met one sort of singularity,the branch point, <strong>and</strong> in this section we will consider other types of singularityas well as discuss the zeros of complex functions.If f(z) has a singular point at z = z 0 but is analytic at all points in someneighbourhood containing z 0 but no other singularities, then z = z 0 is called anisolated singularity. (Clearly, branch points are not isolated singularities.)The most important type of isolated singularity is the pole. Iff(z) has the <strong>for</strong>mf(z) =g(z)(z − z 0 ) n , (24.23)where n is a positive integer, g(z) is analytic at all points in some neighbourhoodcontaining z = z 0 <strong>and</strong> g(z 0 ) ≠0,thenf(z) has a pole of order n at z = z 0 .Analternative (though equivalent) definition is thatlim [(z − z 0 ) n f(z)] = a, (24.24)z→z 0837


COMPLEX VARIABLESwhere a is a finite, non-zero complex number. We note that if the above limit isequal to zero, then z = z 0 is a pole of order less than n, orf(z) is analytic there;if the limit is infinite then the pole is of an order greater than n. It may also beshown that if f(z) has a pole at z = z 0 ,then|f(z)| →∞as z → z 0 from anydirection in the Arg<strong>and</strong> diagram. § If no finite value of n can be found such that(24.24) is satisfied, then z = z 0 is called an essential singularity.◮Find the singularities of the functions(i) f(z) = 11 − z − 1 , (ii) f(z) =tanhz.1+z(i) If we write f(z) asf(z) = 11 − z − 11+z = 2z(1 − z)(1 + z) ,we see immediately from either (24.23) or (24.24) that f(z) has poles of order 1 (or simplepoles) atz =1<strong>and</strong>z = −1.(ii) In this case we writef(z) =tanhz = sinh z exp z − exp(−z)=cosh z exp z +exp(−z) .Thus f(z) has a singularity when exp z = − exp(−z) or, equivalently, whenexp z =exp[i(2n +1)π]exp(−z),where n is any integer. Equating the arguments of the exponentials we find z =(n + 1 )πi, 2<strong>for</strong> integer n.Furthermore, using l’Hôpital’s rule (see chapter 4) we havelimz→(n+ 1 2 )πi{[z − (n +12 )πi]sinhzcosh z}= limz→(n+ 1 2 )πi{[z − (n +12There<strong>for</strong>e, from (24.24), each singularity is a simple pole. ◭})πi]coshz +sinhz=1.sinh zAnother type of singularity exists at points <strong>for</strong> which the value of f(z) takesan indeterminate <strong>for</strong>m such as 0/0 but lim z→z0 f(z) exists <strong>and</strong> is independentof the direction from which z 0 is approached. Such points are called removablesingularities.◮Show that f(z) =(sinz)/z has a removable singularity at z =0.It is clear that f(z) takes the indeterminate <strong>for</strong>m 0/0 atz = 0. However, by exp<strong>and</strong>ingsin z as a power series in z, we findf(z) = 1 (z ···)− z3z 3! + z55! − =1− z23! + z45! − ··· .§ Although perhaps intuitively obvious, this result really requires <strong>for</strong>mal demonstration by analysis.838


24.7 CONFORMAL TRANSFORMATIONSThus lim z→0 f(z) = 1 independently of the way in which z → 0, <strong>and</strong> so f(z) has a removablesingularity at z =0.◭An expression common in mathematics, but which we have so far avoidedusing explicitly in this chapter, is ‘z tends to infinity’. For a real variable suchas |z| or R, ‘tending to infinity’ has a reasonably well defined meaning. For acomplex variable needing a two-dimensional plane to represent it, the meaning isnot intrinsically well defined. However, it is convenient to have a unique meaning<strong>and</strong> this is provided by the following definition: the behaviour of f(z) at infinityis given by that of f(1/ξ) atξ =0,whereξ =1/z.◮Find the behaviour at infinity of (i) f(z) = a + bz −2 , (ii) f(z) = z(1 + z 2 ) <strong>and</strong> (iii)f(z) =expz.(i) f(z) =a + bz −2 : on putting z =1/ξ, f(1/ξ) =a + bξ 2 , which is analytic at ξ =0;thus f is analytic at z = ∞.(ii) f(z) =z(1 + z 2 ): f(1/ξ) =1/ξ +1/ξ 3 ; thus f has a pole of order 3 at z = ∞.(iii) f(z) =expz : f(1/ξ) = ∑ ∞0 (n!)−1 ξ −n ; thus f has an essential singularity at z = ∞. ◭We conclude this section by briefly mentioning the zeros of a complex function.As the name suggests, if f(z 0 )=0thenz = z 0 is called a zero of the functionf(z). Zeros are classified in a similar way to poles, in that iff(z) =(z − z 0 ) n g(z),where n is a positive integer <strong>and</strong> g(z 0 ) ≠0,thenz = z 0 is called a zero of ordern of f(z). If n =1thenz = z 0 is called a simple zero. It may further be shownthat if z = z 0 is a zero of order n of f(z) then it is also a pole of order n of thefunction 1/f(z).We will return in section 24.11 to the classification of zeros <strong>and</strong> poles in termsof their series expansions.24.7 Con<strong>for</strong>mal trans<strong>for</strong>mationsWe now turn our attention to the subject of trans<strong>for</strong>mations, by which we meana change of coordinates from the complex variable z = x + iy to another, sayw = r + is, by means of a prescribed <strong>for</strong>mula:w = g(z) =r(x, y)+is(x, y).Under such a trans<strong>for</strong>mation, or mapping, the Arg<strong>and</strong> diagram <strong>for</strong> the z-variableis trans<strong>for</strong>med into one <strong>for</strong> the w-variable, although the complete z-plane mightbe mapped onto only a part of the w-plane, or onto the whole of the w-plane, oronto some or all of the w-plane covered more than once.We shall consider only those mappings <strong>for</strong> which w <strong>and</strong> z are related by afunction w = g(z) <strong>and</strong> its inverse z = h(w) with both functions analytic, exceptpossibly at a few isolated points; such mappings are called con<strong>for</strong>mal. Their839


COMPLEX VARIABLESyz 1z 2sC ′ 1w 1w 2C 1C 2z 0θ 2 θ 1xw = g(z)φ 2w 0φ 1rC ′ 2Figure 24.3 Two curves C 1 <strong>and</strong> C 2 in the z-plane, which are mapped ontoC 1 ′ <strong>and</strong> C′ 2 in the w-plane.important properties are that, except at points at which g ′ (z), <strong>and</strong> hence h ′ (z), iszero or infinite:(i) continuous lines in the z-plane trans<strong>for</strong>m into continuous lines in thew-plane;(ii) the angle between two intersecting curves in the z-plane equals the anglebetween the corresponding curves in the w-plane;(iii) the magnification, as between the z-plane <strong>and</strong> the w-plane, of a small lineelement in the neighbourhood of any particular point is independent ofthe direction of the element;(iv) any analytic function of z trans<strong>for</strong>ms to an analytic function of w <strong>and</strong>vice versa.Result (i) is immediate, <strong>and</strong> results (ii) <strong>and</strong> (iii) can be justified by the followingargument. Let two curves C 1 <strong>and</strong> C 2 pass through the point z 0 in the z-plane<strong>and</strong> let z 1 <strong>and</strong> z 2 be two points on their respective tangents at z 0 , each a distanceρ from z 0 . The same prescription with w replacing z describes the trans<strong>for</strong>medsituation; however, the trans<strong>for</strong>med tangents may not be straight lines <strong>and</strong> thedistances of w 1 <strong>and</strong> w 2 from w 0 have not yet been shown to be equal. Thissituation is illustrated in figure 24.3.In the z-plane z 1 <strong>and</strong> z 2 are given byz 1 − z 0 = ρ exp iθ 1 <strong>and</strong> z 2 − z 0 = ρ exp iθ 2 .The corresponding descriptions in the w-plane arew 1 − w 0 = ρ 1 exp iφ 1 <strong>and</strong> w 2 − w 0 = ρ 2 exp iφ 2 .The angles θ i <strong>and</strong> φ i are clear from figure 24.3. The trans<strong>for</strong>med angles φ i arethose made with the r-axis by the tangents to the trans<strong>for</strong>med curves at their840


24.7 CONFORMAL TRANSFORMATIONSpoint of intersection. Since any finite-length tangent may be curved, w i is morestrictly given by w i − w 0 = ρ i exp i(φ i + δφ i ), where δφ i → 0asρ i → 0, i.e. asρ → 0.Now since w = g(z), where g is analytic, we have( ) ( )w1 − w 0w2 − w 0lim= lim= dgz 1→z 0 z 1 − z 0 z2→z 0 z 2 − z 0 dz ∣ ,z=z0which may be written as{ρ1limρ→0 ρ exp[i(φ 1 + δφ 1 − θ 1 )]}{ }ρ2= limρ→0 ρ exp[i(φ 2 + δφ 2 − θ 2 )] = g ′ (z 0 ).(24.25)Comparing magnitudes <strong>and</strong> phases (i.e. arguments) in the equalities (24.25) givesthe stated results (ii) <strong>and</strong> (iii) <strong>and</strong> adds quantitative in<strong>for</strong>mation to them, namelythat <strong>for</strong> small line elementsρ 1ρ ≈ ρ 2ρ ≈|g′ (z 0 )|, (24.26)φ 1 − θ 1 ≈ φ 2 − θ 2 ≈ arg g ′ (z 0 ). (24.27)For strict comparison with result (ii), (24.27) must be written as θ 1 −θ 2 = φ 1 −φ 2 ,with an ordinary equality sign, since the angles are only defined in the limitρ → 0 when (24.27) becomes a true identity. We also see from (24.26) thatthe linear magnification factor is |g ′ (z 0 )|; similarly, small areas are magnified by|g ′ (z 0 )| 2 .Since in the neighbourhoods of corresponding points in a trans<strong>for</strong>mation anglesare preserved <strong>and</strong> magnifications are independent of direction, it follows that smallplane figures are trans<strong>for</strong>med into figures of the same shape, but, in general, onesthat are magnified <strong>and</strong> rotated (though not distorted). However, we also notethat at points where g ′ (z) = 0, the angle arg g ′ (z) through which line elements arerotated is undefined; these are called critical points of the trans<strong>for</strong>mation.The final result (iv) is perhaps the most important property of con<strong>for</strong>maltrans<strong>for</strong>mations. If f(z) is an analytic function of z <strong>and</strong> z = h(w) is also analytic,then F(w) =f(h(w)) is analytic in w. Its importance lies in the further conclusionsit allows us to draw from the fact that, since f is analytic, the real <strong>and</strong> imaginaryparts of f = φ + iψ are necessarily solutions of∂ 2 φ∂x 2 + ∂2 φ∂y 2 = 0 <strong>and</strong> ∂ 2 ψ∂x 2 + ∂2 ψ=0. (24.28)∂y2 Since the trans<strong>for</strong>mation property ensures that F =Φ+iΨ is also analytic, wecan conclude that its real <strong>and</strong> imaginary parts must themselves satisfy Laplace’sequation in the w-plane:∂ 2 Φ∂r 2+ ∂2 Φ∂s 2 = 0 <strong>and</strong> ∂ 2 Ψ∂r 2841+ ∂2 Ψ=0. (24.29)∂s2


COMPLEX VARIABLESysi Pw = g(z)Q R S TxR ′ P ′ Q ′T ′S ′rFigure 24.4 Trans<strong>for</strong>ming the upper half of the z-plane into the interior ofthe unit circle in the w-plane, in such a way that z = i is mapped onto w =0<strong>and</strong> the points x = ±∞ are mapped onto w =1.Further, suppose that (say) Re f(z) =φ is constant over a boundary C in thez-plane; then Re F(w) = Φ is constant over C in the z-plane. But this is the sameas saying that Re F(w) is constant over the boundary C ′ in the w-plane, C ′ beingthe curve into which C is trans<strong>for</strong>med by the con<strong>for</strong>mal trans<strong>for</strong>mation w = g(z).This result is exploited extensively in the next chapter to solve Laplace’s equation<strong>for</strong> a variety of two-dimensional geometries.Examples of useful con<strong>for</strong>mal trans<strong>for</strong>mations are numerous. For instance,w = z + b, w =(expiφ)z <strong>and</strong> w = az correspond, respectively, to a translation byb, a rotation through an angle φ <strong>and</strong> a stretching (or contraction) in the radialdirection (<strong>for</strong> a real). These three examples can be combined into the generallinear trans<strong>for</strong>mation w = az +b, where,ingeneral,a <strong>and</strong> b are complex. Anotherexample is the inversion mapping w =1/z, which maps the interior of the unitcircle to the exterior <strong>and</strong> vice versa. Other, more complicated, examples also exist.◮Show that if the point z 0 lies in the upper half of the z-plane then the trans<strong>for</strong>mationw =(expiφ) z − z 0z − z0∗maps the upper half of the z-plane into the interior of the unit circle in the w-plane. Hencefind a similar trans<strong>for</strong>mation that maps the point z = i onto w =0<strong>and</strong> the points x = ±∞onto w =1.Taking the modulus of w, we have|w| =∣ (exp iφ) z − z ∣ 0∣∣∣ z − z0∗ ∣ = z − z 0z − z0∗ ∣ .However, since the complex conjugate z0 ∗ is the reflection of z 0 in the real axis, if z <strong>and</strong> z 0both lie in the upper half of the z-plane then |z − z 0 |≤|z − z0 ∗ |; thus |w| ≤1, as required.We also note that (i) the equality holds only when z lies on the real axis, <strong>and</strong> so this axisis mapped onto the boundary of the unit circle in the w-plane; (ii) the point z 0 is mappedonto w = 0, the origin of the w-plane.By fixing the images of two points in the z-plane, the constants z 0 <strong>and</strong> φ can also befixed. Since we require the point z = i to be mapped onto w = 0, we have immediately842


24.7 CONFORMAL TRANSFORMATIONSysw 5w 1w 4w = g(z) φφ 41x 1 x 2 x 3 x 4 x 5xrφ 2 φ 3w 2w 3φ 5Figure 24.5 Trans<strong>for</strong>ming the upper half of the z-plane into the interiorof a polygon in the w-plane, in such a way that the points x 1 ,x 2 ,...,x n aremapped onto the vertices w 1 ,w 2 ,...,w n of the polygon with interior anglesφ 1 ,φ 2 ,...,φ n .z 0 = i. By further requiring z = ±∞ to be mapped onto w = 1, we find 1 = w =expiφ<strong>and</strong> so φ = 0. The required trans<strong>for</strong>mation is there<strong>for</strong>ew = z − iz + i ,<strong>and</strong> is illustrated in figure 24.4. ◭We conclude this section by mentioning the rather curious Schwarz–Christoffeltrans<strong>for</strong>mation. § Suppose, as shown in figure 24.5, that we are interested in a(finite) number of points x 1 ,x 2 ,...,x n on the real axis in the z-plane. Then bymeans of the trans<strong>for</strong>mation{ ∫ z}w = A (ξ − x 1 ) (φ1/π)−1 (ξ − x 2 ) (φ2/π)−1 ···(ξ − x n ) (φn/π)−1 dξ + B, (24.30)0we may map the upper half of the z-plane onto the interior of a closed polygon inthe w-plane having n vertices w 1 ,w 2 ,...,w n (which are the images of x 1 ,x 2 ,...,x n )with corresponding interior angles φ 1 ,φ 2 ,...,φ n , as shown in figure 24.5. Thereal axis in the z-plane is trans<strong>for</strong>med into the boundary of the polygon itself.The constants A <strong>and</strong> B are complex in general <strong>and</strong> determine the position,size <strong>and</strong> orientation of the polygon. It is clear from (24.30) that dw/dz =0atx = x 1 ,x 2 ,...,x n , <strong>and</strong> so the trans<strong>for</strong>mation is not con<strong>for</strong>mal at these points.There are various subtleties associated with the use of the Schwarz–Christoffeltrans<strong>for</strong>mation. For example, if one of the points on the real axis in the z-plane(usually x n ) is taken at infinity, then the corresponding factor in (24.30) (i.e. theone involving x n ) is not present. In this case, the point(s) x = ±∞ are consideredas one point, since they trans<strong>for</strong>m to a single vertex of the polygon in the w-plane.§ Strictly speaking, the use of this trans<strong>for</strong>mation requires an underst<strong>and</strong>ing of complex integrals,which are discussed in section 24.8.843


COMPLEX VARIABLESysibw 3w = g(z)φ 3x 1 x 2−1 1xw 1φ 1 φ 2 w 2−aarFigure 24.6 Trans<strong>for</strong>ming the upper half of the z-plane into the interior ofa triangle in the w-plane.We can also map the upper half of the z-plane into an infinite open polygonby considering it as the limiting case of some closed polygon.◮Find a trans<strong>for</strong>mation that maps the upper half of the z-plane into the triangular regionshown in figure 24.6 in such a way that the points x 1 = −1 <strong>and</strong> x 2 = 1 are mappedinto the points w = −a <strong>and</strong> w = a, respectively, <strong>and</strong> the point x 3 = ±∞ is mapped intow = ib. Hence find a trans<strong>for</strong>mation that maps the upper half of the z-plane into the region−a


24.8 COMPLEX INTEGRALSyw 3 s w 3w = g(z)x1 x 2−1 1xφ 1 φ 2w 1 w 2−aarFigure 24.7 Trans<strong>for</strong>ming the upper half of the z-plane into the interior ofthe region −a


COMPLEX VARIABLESyBC 2C 1xC 3AFigure 24.8A <strong>and</strong> B.Some alternative paths <strong>for</strong> the integral of a function f(z) betweenThe question of when such an integral exists will not be pursued, except to statethat a sufficient condition is that dx/dt <strong>and</strong> dy/dt are continuous.◮Evaluate the complex integral of f(z) =z −1 along the circle |z| = R, starting <strong>and</strong> finishingat z = R.The path C 1 is parameterised as follows (figure 24.9(a)):z(t) =R cos t + iR sin t, 0 ≤ t ≤ 2π,whilst f(z) isgivenbyf(z) = 1x + iy = x − iyx 2 + y . 2Thus the real <strong>and</strong> imaginary parts of f(z) areu =xx 2 + y = R cos t <strong>and</strong> v = −y2 R 2 x 2 + y = − R sin t .2 R 2Hence, using expression (24.34),∫ ∫12π∫C 1z dz = cos t2π( − sin t0 R(−R sin t) dt − 0 R∫ 2π∫cos t2π( − sin t+ i0 RR cos tdt+ i 0 R=0+0+iπ + iπ =2πi. ◭)R cos tdt)(−R sin t) dt (24.35)With a bit of experience, the reader may be able to evaluate integrals likethe LHS of (24.35) directly without having to write them as four separate realintegrals. In the present case,∫C 1∫dz 2πz = −R sin t + iR cos tdt =0 R cos t + iR sin t846∫ 2π0idt=2πi. (24.36)


24.8 COMPLEX INTEGRALSyRtyyC 1 C 2Rt s =1t =0x −RR x −RR xC 3biRC3a(a) (b) (c)Figure 24.9details.Different paths <strong>for</strong> an integral of f(z) =z −1 . See the text <strong>for</strong>This very important result will be used many times later, <strong>and</strong> the following shouldbe carefully noted: (i) its value, (ii) that this value is independent of R.In the above example the contour was closed, <strong>and</strong> so it began <strong>and</strong> ended atthe same point in the Arg<strong>and</strong> diagram. We can evaluate complex integrals alongopen paths in a similar way.◮Evaluate the complex integral of f(z) =z −1 along the following paths (see figure 24.9):(i) the contour C 2 consisting of the semicircle |z| = R in the half-plane y ≥ 0,(ii) the contour C 3 made up of the two straight lines C 3a <strong>and</strong> C 3b .(i) This is just as in the previous example, except that now 0 ≤ t ≤ π. With this change,we have from (24.35) or (24.36) that∫dz= πi. (24.37)C 2z(ii) The straight lines that make up the countour C 3 may be parameterised as follows:C 3a , z =(1− t)R + itR <strong>for</strong> 0 ≤ t ≤ 1;C 3b , z = −sR + i(1 − s)R <strong>for</strong> 0 ≤ s ≤ 1.With these parameterisations the required integrals may be written∫ ∫dz 1∫C 3z = −R + iR1R + t(−R + iR) dt + −R − iRds. (24.38)iR + s(−R − iR)0If we could take over from real-variable theory that, <strong>for</strong> real t, ∫ (a+bt) −1 dt = b −1 ln(a+bt)even if a <strong>and</strong> b are complex, then these integrals could be evaluated immediately. However,to do this would be presuming to some extent what we wish to show, <strong>and</strong> so the evaluation8470


COMPLEX VARIABLESmust be made in terms of entirely real integrals. For example, the first is given by∫ 10∫−R + iR1R(1 − t)+itR dt =0∫ 1(−1+i)(1 − t − it)(1 − t) 2 + t 2 dt∫2t − 11=0 1 − 2t +2t dt + i 12 0 1 − 2t +2t dt 2= 1 [] 1[ (ln(1 − 2t +2t 2 ) + i t −12tan −1 212022[ π(2 − − π )]= πi2 2 .=0+ i 2The second integral on the RHS of (24.38) can also be shown to have the value πi/2. Thus∫dzC 3z = πi. ◭Considering the results of the preceding two examples, which have commonintegr<strong>and</strong>s <strong>and</strong> limits, some interesting observations are possible. Firstly, the twointegrals from z = R to z = −R, along C 2 <strong>and</strong> C 3 , respectively, have the samevalue, even though the paths taken are different. It also follows that if we took aclosed path C 4 , given by C 2 from R to −R <strong>and</strong> C 3 traversed backwards from −Rto R, then the integral round C 4 of z −1 would be zero (both parts contributingequal <strong>and</strong> opposite amounts). This is to be compared with result (24.36), in whichclosed path C 1 , beginning <strong>and</strong> ending at the same place as C 4 , yields a value 2πi.It is not true, however, that the integrals along the paths C 2 <strong>and</strong> C 3 are equal<strong>for</strong> any function f(z), or, indeed, that their values are independent of R in general.◮Evaluate the complex integral of f(z) =Rez along the paths C 1 , C 2 <strong>and</strong> C 3 shown infigure 24.9.(i) If we take f(z) =Rez <strong>and</strong> the contour C 1 then∫∫ 2πRe zdz= R cos t(−R sin t + iR cos t) dt = iπR 2 .C 1 0(ii) Using C 2 as the contour,∫∫ πRe zdz= R cos t(−R sin t + iR cos t) dt = 1 2 iπR2 .C 2 0(iii) Finally the integral along C 3 = C 3a + C 3b is given by∫∫ 1∫ 1Re zdz= (1 − t)R(−R + iR) dt + (−sR)(−R − iR) dsC 3 00= 1 2 R2 (−1+i)+ 1 2 R2 (1 + i) =iR 2 . ◭)] 10The results of this section demonstrate that the value of an integral between thesame two points may depend upon the path that is taken between them but, atthe same time, suggest that, under some circumstances, the value is independentof the path. The general situation is summarised in the result of the next section,848


24.9 CAUCHY’S THEOREMnamely Cauchy’s theorem, which is the cornerstone of the integral calculus ofcomplex variables.Be<strong>for</strong>e discussing Cauchy’s theorem, however, we note an important resultconcerning complex integrals that will be of some use later. Let us consider theintegral of a function f(z) along some path C. IfM is an upper bound on thevalue of |f(z)| on the path, i.e. |f(z)| ≤M on C, <strong>and</strong>L is the length of the path C,then∫∫∫∣ f(z) dz∣ ≤ |f(z)||dz| ≤M dl = ML. (24.39)CcCIt is straight<strong>for</strong>ward to verify that this result does indeed hold <strong>for</strong> the complexintegrals considered earlier in this section.24.9 Cauchy’s theoremCauchy’s theorem states that if f(z) is an analytic function, <strong>and</strong> f ′ (z) is continuousat each point within <strong>and</strong> on a closed contour C, then∮f(z) dz =0. (24.40)CIn this statement <strong>and</strong> from now on we denote an integral around a closed contourby ∮ C .To prove this theorem we will need the two-dimensional <strong>for</strong>m of the divergencetheorem, known as Green’s theorem in a plane (see section 11.3). This says thatif p <strong>and</strong> q are two functions with continuous first derivatives within <strong>and</strong> on aclosed contour C (bounding a domain R) inthexy-plane, then∫∫R( ∂p∂x + ∂q∂y)dxdy =∮C(p dy− qdx). (24.41)With f(z) =u + iv <strong>and</strong> dz = dx + idy, this can be applied to∮∮∮I = f(z) dz = (u dx− vdy)+i (v dx+ udy)to give∫∫ [ ∂(−u)I =∂yRCC+ ∂(−v) ] ∫∫ [ ∂(−v)dx dy + i+ ∂u ]dx dy. (24.42)∂xR ∂y ∂xNow, recalling that f(z) is analytic <strong>and</strong> there<strong>for</strong>e that the Cauchy–Riemannrelations (24.5) apply, we see that each integr<strong>and</strong> is identically zero <strong>and</strong> thus I isalso zero; this proves Cauchy’s theorem.In fact, the conditions of the above proof are more stringent than they needbe. The continuity of f ′ (z) is not necessary <strong>for</strong> the proof of Cauchy’s theorem,849C


COMPLEX VARIABLESyBC 1RC 2AxFigure 24.10 Two paths C 1 <strong>and</strong> C 2 enclosing a region R.analyticity of f(z) within <strong>and</strong> on C being sufficient. However, the proof thenbecomes more complicated <strong>and</strong> is too long to be given here. §The connection between Cauchy’s theorem <strong>and</strong> the zero value of the integralof z −1 around the composite path C 4 discussed towards the end of the previoussection is apparent: the function z −1 is analytic in the two regions of the z-planeenclosed by contours (C 2 <strong>and</strong> C 3a )<strong>and</strong>(C 2 <strong>and</strong> C 3b ).◮Suppose two points A <strong>and</strong> B in the complex plane are joined by two different paths C 1<strong>and</strong> C 2 . Show that if f(z) is an analytic function on each path <strong>and</strong> in the region enclosedby the two paths, then the integral of f(z) is the same along C 1 <strong>and</strong> C 2 .The situation is shown in figure 24.10. Since f(z) isanalyticinR, it follows from Cauchy’stheorem that we have∫ ∫∮f(z) dz − f(z) dz = f(z) dz =0,C 1 C 2 C 1 −C 2since C 1 − C 2 <strong>for</strong>ms a closed contour enclosing R. Thus we immediately obtain∫∫f(z) dz = f(z) dz,C 1 C 2<strong>and</strong> so the values of the integrals along C 1 <strong>and</strong> C 2 are equal. ◭An important application of Cauchy’s theorem is in proving that, in somecases, it is possible to de<strong>for</strong>m a closed contour C into another contour γ in sucha way that the integrals of a function f(z) around each of the contours have thesame value.§ The reader may refer to almost any book that is devoted to complex variables <strong>and</strong> the theory offunctions.850


24.10 CAUCHY’S INTEGRAL FORMULAyCC 1γC 2Figure 24.11 The contour used to prove the result (24.43).x◮Consider two closed contours C <strong>and</strong> γ in the Arg<strong>and</strong> diagram, γ being sufficiently smallthat it lies completely within C. Show that if the function f(z) is analytic in the regionbetween the two contours then∮∮f(z) dz = f(z) dz. (24.43)CγTo prove this result we consider a contour as shown in figure 24.11. The two close parallellines C 1 <strong>and</strong> C 2 join γ <strong>and</strong> C, which are ‘cut’ to accommodate them. The new contour Γso <strong>for</strong>med consists of C, C 1 , γ <strong>and</strong> C 2 .Within the area bounded by Γ, the function f(z) is analytic, <strong>and</strong> there<strong>for</strong>e, by Cauchy’stheorem (24.40),∮f(z) dz =0. (24.44)ΓNow the parts C 1 <strong>and</strong> C 2 of Γ are traversed in opposite directions, <strong>and</strong> in the limit lie ontop of each other, <strong>and</strong> so their contributions to (24.44) cancel. Thus∮ ∮f(z) dz + f(z) dz =0. (24.45)CThe sense of the integral round γ is opposite to the conventional (anticlockwise) one, <strong>and</strong>so by traversing γ in the usual sense, we establish the result (24.43). ◭A sort of converse of Cauchy’s theorem is known as Morera’s theorem, whichstates that if f(z) is a continuous function of z in a closed domain R bounded byacurveC <strong>and</strong>, further, ∮ Cf(z) dz =0,thenf(z) is analytic in R.γ24.10 Cauchy’s integral <strong>for</strong>mulaAnother very important theorem in the theory of complex variables is Cauchy’sintegral <strong>for</strong>mula, which states that if f(z) is analytic within <strong>and</strong> on a closed851


COMPLEX VARIABLEScontour C <strong>and</strong> z 0 is a point within C thenf(z 0 )= 1 ∮f(z)dz. (24.46)2πi C z − z 0This <strong>for</strong>mula is saying that the value of an analytic function anywhere inside aclosed contour is uniquely determined by its values on the contour § <strong>and</strong> that thespecific expression (24.46) can be given <strong>for</strong> the value at the interior point.We may prove Cauchy’s integral <strong>for</strong>mula by using (24.43) <strong>and</strong> taking γ to be acircle centred on the point z = z 0 , of small enough radius ρ that it all lies insideC. Then, since f(z) is analytic inside C, the integr<strong>and</strong> f(z)/(z − z 0 ) is analyticin the space between C <strong>and</strong> γ. Thus, from (24.43), the integral around γ has thesame value as that around C.We then use the fact that any point z on γ is given by z = z 0 + ρ exp iθ (<strong>and</strong>so dz = iρ exp iθ dθ). Thus the value of the integral around γ is given by∮I =γf(z)z − z 0dz == i∫ 2π0∫ 2π0f(z 0 + ρ exp iθ)iρ exp iθ dθρ exp iθf(z 0 + ρ exp iθ) dθ.If the radius of the circle γ is now shrunk to zero, i.e. ρ → 0, then I → 2πif(z 0 ),thus establishing the result (24.46).An extension to Cauchy’s integral <strong>for</strong>mula can be made, yielding an integralexpression <strong>for</strong> f ′ (z 0 ):f ′ (z 0 )= 1 ∫2πi Cunder the same conditions as previously stated.◮Prove Cauchy’s integral <strong>for</strong>mula <strong>for</strong> f ′ (z 0 ) given in (24.47).f(z)dz, (24.47)(z − z 0 )2To show this, we use the definition of a derivative <strong>and</strong> (24.46) itself to evaluatef ′ f(z 0 + h) − f(z 0 )(z 0 ) = limh→0 h[ ∮ (1 f(z) 1= limh→0 2πi C h z − z 0 − h − 1 ) ]dzz − z 0[ ∮]1f(z)= limh→0 2πi (z − z 0 − h)(z − z 0 ) dz= 12πiwhich establishes result (24.47). ◭∮CCf(z)(z − z 0 ) 2 dz,§ The similarity between this <strong>and</strong> the uniqueness theorem <strong>for</strong> the Laplace equation with Dirichletboundary conditions (see chapter 20) is apparent.852


24.11 TAYLOR AND LAURENT SERIESFurther, it may be proved by induction that the nth derivative of f(z) is alsogiven by a Cauchy integral,f (n) (z 0 )= n! ∮f(z) dz. (24.48)2πi C (z − z 0 )n+1Thus, if the value of the analytic function is known on C then not only may thevalue of the function at any interior point be calculated, but also the values ofall its derivatives.The observant reader will notice that (24.48) may also be obtained by the<strong>for</strong>mal device of differentiating under the integral sign with respect to z 0 inCauchy’s integral <strong>for</strong>mula (24.46):f (n) (z 0 )= 1 ∮∂ n [ ] f(z)dz2πi (z − z 0 )= n!2πi∮C ∂z0nCf(z) dz(z − z 0 ) n+1 .◮Suppose that f(z) is analytic inside <strong>and</strong> on a circle C of radius R centred on the pointz = z 0 .If|f(z)| ≤M on the circle, where M is some constant, show that|f (n) (z 0 )|≤ Mn!R . (24.49)nFrom (24.48) we have|f (n) (z 0 )| = n!∮2π ∣<strong>and</strong> on using (24.39) this becomes|f (n) (z 0 )|≤ n! M2π RThis result is known as Cauchy’s inequality. ◭Cf(z) dz(z − z 0 ) n+1 ∣ ∣∣∣,Mn!2πR =n+1R . nWe may use Cauchy’s inequality to prove Liouville’s theorem, which states thatif f(z) is analytic <strong>and</strong> bounded <strong>for</strong> all z then f is a constant. Setting n =1in(24.49) <strong>and</strong> letting R →∞, we find |f ′ (z 0 )| =0<strong>and</strong>hencef ′ (z 0 ) = 0. Since f(z) isanalytic <strong>for</strong> all z, wemaytakez 0 as any point in the z-plane <strong>and</strong> thus f ′ (z) =0<strong>for</strong> all z; this implies f(z) = constant. Liouville’s theorem may be used in turn toprove the fundamental theorem of algebra (see exercise 24.9).24.11 Taylor <strong>and</strong> Laurent seriesFollowing on from (24.48), we may establish Taylor’s theorem <strong>for</strong>functionsofacomplex variable. If f(z) is analytic inside <strong>and</strong> on a circle C of radius R centredon the point z = z 0 ,<strong>and</strong>z is a point inside C, then∞∑f(z) = a n (z − z 0 ) n , (24.50)n=0853


COMPLEX VARIABLESwhere a n is given by f (n) (z 0 )/n!. The Taylor expansion is valid inside the regionof analyticity <strong>and</strong>, <strong>for</strong> any particular z 0 , can be shown to be unique.To prove Taylor’s theorem (24.50), we note that, since f(z) is analytic inside<strong>and</strong> on C, we may use Cauchy’s <strong>for</strong>mula to write f(z) asf(z) = 12πi∮Cn=0f(ξ)dξ, (24.51)ξ − zwhere ξ lies on C. Now we may exp<strong>and</strong> the factor (ξ − z) −1 as a geometric seriesin (z − z 0 )/(ξ − z 0 ),1ξ − z = 1 ∑∞ ( ) n z − z0,ξ − z 0 ξ − z 0so (24.51) becomesf(z) = 1 ∮f(ξ)2πi C ξ − z 0= 1 ∞∑2πi= 12πin=0∞ ∑n=0(z − z 0 ) n ∮( z − z0ξ − z 0) nCdξf(ξ)dξ(ξ − z 0 )n+1∞∑(z − z 0 ) n 2πif(n) (z 0 ), (24.52)n!n=0where we have used Cauchy’s integral <strong>for</strong>mula (24.48) <strong>for</strong> the derivatives off(z). Cancelling the factors of 2πi, we thus establish the result (24.50) witha n = f (n) (z 0 )/n!.◮Show that if f(z) <strong>and</strong> g(z) are analytic in some region R, <strong>and</strong> f(z) =g(z) within somesubregion S of R, thenf(z) =g(z) throughout R.It is simpler to consider the (analytic) function h(z) =f(z)−g(z), <strong>and</strong> to show that becauseh(z) =0inS it follows that h(z) = 0 throughout R.If we choose a point z = z 0 in S, then we can exp<strong>and</strong> h(z) in a Taylor series about z 0 ,h(z) =h(z 0 )+h ′ (z 0 )(z − z 0 )+ 1 2 h′′ (z 0 )(z − z 0 ) 2 + ··· ,which will converge inside some circle C that extends at least as far as the nearest part ofthe boundary of R, sinceh(z) isanalyticinR. But since z 0 lies in S, we haveh(z 0 )=h ′ (z 0 )=h ′′ (z 0 )=···=0,<strong>and</strong> so h(z) =0insideC. We may now exp<strong>and</strong> about a new point, which can lie anywherewithin C, <strong>and</strong> repeat the process. By continuing this procedure we may show that h(z) =0throughout R.This result is called the identity theorem <strong>and</strong>, in fact, the equality of f(z) <strong>and</strong>g(z)throughout R follows from their equality along any curve of non-zero length in R, orevenat a countably infinite number of points in R. ◭So far we have assumed that f(z) is analytic inside <strong>and</strong> on the (circular)contour C. If,however,f(z) has a singularity inside C at the point z = z 0 ,thenitcannot be exp<strong>and</strong>ed in a Taylor series. Nevertheless, suppose that f(z) has a pole854


24.11 TAYLOR AND LAURENT SERIESof order p at z = z 0 but is analytic at every other point inside <strong>and</strong> on C. Thenthe function g(z) =(z − z 0 ) p f(z) is analytic at z = z 0 , <strong>and</strong> so may be exp<strong>and</strong>edas a Taylor series about z = z 0 :∞∑g(z) = b n (z − z 0 ) n . (24.53)n=0Thus, <strong>for</strong> all z inside C, f(z) will have a power series representation of the <strong>for</strong>mf(z) =a −p(z − z 0 ) p + ···+ a −1+ a 0 + a 1 (z − z 0 )+a 2 (z − z 0 ) 2 + ··· ,z − z 0(24.54)with a −p ≠ 0. Such a series, which is an extension of the Taylor expansion, iscalled a Laurent series. By comparing the coefficients in (24.53) <strong>and</strong> (24.54), wesee that a n = b n+p . Now, the coefficients b n in the Taylor expansion of g(z) areseen from (24.52) to be given byb n = g(n) (z 0 )= 1 ∮g(z)dz,n! 2πi (z − z 0 )n+1<strong>and</strong> so <strong>for</strong> the coefficients a n in (24.54) we havea n = 12πi∮g(z)1dz =(z − z 0 )n+1+p2πi∮f(z)dz,(z − z 0 )n+1an expression that is valid <strong>for</strong> both positive <strong>and</strong> negative n.The terms in the Laurent series with n ≥ 0 are collectively called the analyticpart, whilst the remainder of the series, consisting of terms in inverse powers ofz − z 0 , is called the principal part. Depending on the nature of the point z = z 0 ,the principal part may contain an infinite number of terms, so thatf(z) =+∞∑n=−∞a n (z − z 0 ) n . (24.55)In this case we would expect the principal part to converge only <strong>for</strong> |(z − z 0 ) −1 |less than some constant, i.e. outside some circle centred on z 0 . However, theanalytic part will converge inside some (different) circle also centred on z 0 .Ifthelatter circle has the greater radius then the Laurent series will converge in theregion R between the two circles (see figure 24.12); otherwise it does not convergeat all.In fact, it may be shown that any function f(z) that is analytic in a regionR between two such circles C 1 <strong>and</strong> C 2 centred on z = z 0 can be expressed asa Laurent series about z 0 that converges in R. We note that, depending on thenature of the point z = z 0 , the inner circle may be a point (when the principalpart contains only a finite number of terms) <strong>and</strong> the outer circle may have aninfinite radius.We may use the Laurent series of a function f(z) about any point z = z 0 to855


COMPLEX VARIABLESyRz 0C 1C 2xFigure 24.12 The region of convergence R <strong>for</strong> a Laurent series of f(z) abouta point z = z 0 where f(z) has a singularity.classify the nature of that point. If f(z) is actually analytic at z = z 0 ,thenin(24.55) all a n <strong>for</strong> n0;(ii) it is not possible to find such a lowest value of −p.In case (i), f(z) is of the <strong>for</strong>m (24.54) <strong>and</strong> is described as having a pole of orderp at z = z 0 ; the value of a −1 (not a −p ) is called the residue of f(z) at the polez = z 0 , <strong>and</strong> will play an important part in later applications.For case (ii), in which the negatively decreasing powers of z − z 0 do notterminate, f(z) is said to have an essential singularity. These definitions should becompared with those given in section 24.6.◮Find the Laurent series off(z) =1z(z − 2) 3about the singularities z =0<strong>and</strong> z =2(separately). Hence verify that z =0is a pole o<strong>for</strong>der 1 <strong>and</strong> z =2is a pole of order 3, <strong>and</strong> find the residue of f(z) at each pole.To obtain the Laurent series about z = 0, we make the factor in parentheses in the856


24.11 TAYLOR AND LAURENT SERIESdenominator take the <strong>for</strong>m (1 − αz), where α is some constant, <strong>and</strong> thus obtain1f(z)=−8z(1 − z/2) 3=− 1 [ (1+(−3) − z )+ (−3)(−4) (− z ) 2 (−3)(−4)(−5)(+ − z ) 3+ ···]8z2 2! 2 3! 2= − 1 8z − 316 − 3z16 − 5z232 − ··· .Since the lowest power of z is −1, the point z = 0 is a pole of order 1. The residue of f(z)at z = 0 is simply the coefficient of z −1 in the Laurent expansion about that point <strong>and</strong> isequal to −1/8.The Laurent series about z = 2 is most easily found by letting z =2+ξ (or z − 2=ξ)<strong>and</strong> substituting into the expression <strong>for</strong> f(z) toobtainf(z) =1(2 + ξ)ξ 3 = 1= 12ξ 3 [1 −2ξ 3 (1 + ξ/2)( ) 2 ( ξ ξ−2 2( ξ2)+) 3 ( ) 4 ξ+ − ···]2= 12ξ − 13 4ξ + 12 8ξ − 1 16 + ξ 32 − ···1=2(z − 2) − 13 4(z − 2) + 12 8(z − 2) − 116 + z − 232 − ··· .From this series we see that z = 2 is a pole of order 3 <strong>and</strong> that the residue of f(z) atz =2is 1/8. ◭As we shall see in the next few sections, finding the residue of a functionat a singularity is of crucial importance in the evaluation of complex integrals.Specifically, <strong>for</strong>mulae exist <strong>for</strong> calculating the residue of a function at a particular(singular) point z = z 0 without having to exp<strong>and</strong> the function explicitly as aLaurent series about z 0 <strong>and</strong> identify the coefficient of (z − z 0 ) −1 . The type of<strong>for</strong>mula generally depends on the nature of the singularity at which the residueis required.◮Suppose that f(z) has a pole of order m at the point z = z 0 . By considering the Laurentseries of f(z) about z 0 , derive a general expression <strong>for</strong> the residue R(z 0 ) of f(z) at z = z 0 .Hence evaluate the residue of the functionat the point z = i.f(z) =exp iz(z 2 +1) 2If f(z) has a pole of order m at z = z 0 , then its Laurent series about this point has the<strong>for</strong>mf(z) =a −m(z − z 0 ) + ···+ a −1m (z − z 0 ) + a 0 + a 1 (z − z 0 )+a 2 (z − z 0 ) 2 + ··· ,which, on multiplying both sides of the equation by (z − z 0 ) m ,gives(z − z 0 ) m f(z) =a −m + a −m+1 (z − z 0 )+···+ a −1 (z − z 0 ) m−1 + ··· .857


COMPLEX VARIABLESDifferentiating both sides m − 1times,weobtaind m−1∞∑dz [(z − z 0) m f(z)] = (m − 1)! a m−1 −1 + b n (z − z 0 ) n ,n=1<strong>for</strong> some coefficients b n . In the limit z → z 0 , however, the terms in the sum disappear, <strong>and</strong>after rearranging we obtain the <strong>for</strong>mula{}1 d m−1R(z 0 )=a −1 = limz→z0 (m − 1)! dz [(z − z 0) m f(z)] , (24.56)m−1which gives the value of the residue of f(z) at the point z = z 0 .If we now consider the functionexp izf(z) =(z 2 +1) = exp iz2 (z + i) 2 (z − i) , 2we see immediately that it has poles of order 2 (double poles) at z = i <strong>and</strong> z = −i. Tocalculate the residue at (<strong>for</strong> example) z = i, we may apply the <strong>for</strong>mula (24.56) with m =2.Per<strong>for</strong>ming the required differentiation, we obtainddz [(z − i)2 f(z)] = d [ ] exp izdz (z + i) 21=(z + i) [(z + 4 i)2 i exp iz − 2(exp iz)(z + i)].Setting z = i, we find the residue is given byR(i) = 1 1 (−4ie −1 − 4ie −1) = − i1! 162e . ◭An important special case of (24.56) occurs when f(z) has a simple pole (a poleof order 1) at z = z 0 . Then the residue at z 0 is given byR(z 0 ) = lim [(z − z 0 )f(z)] . (24.57)z→z0If f(z) has a simple pole at z = z 0 <strong>and</strong>, as is often the case, has the <strong>for</strong>mg(z)/h(z), where g(z) is analytic <strong>and</strong> non-zero at z 0 <strong>and</strong> h(z 0 ) = 0, then (24.57)becomes(z − z 0 )g(z)(z − z 0 )R(z 0 ) = lim= g(z 0 ) limz→z0 h(z)z→z0 h(z)1= g(z 0 ) limz→z0 h ′ (z) = g(z 0)h ′ (z 0 ) , (24.58)where we have used l’Hôpital’s rule. This result often provides the simplest wayof determining the residue at a simple pole.24.12 Residue theoremHaving seen from Cauchy’s theorem that the value of an integral round a closedcontour C is zero if the integr<strong>and</strong> is analytic inside the contour, it is natural toask what value it takes when the integr<strong>and</strong> is not analytic inside C. Theanswerto this is contained in the residue theorem, which we now discuss.858


24.12 RESIDUE THEOREMSuppose the function f(z) has a pole of order m at the point z = z 0 ,<strong>and</strong>socan be written as a Laurent series about z 0 of the <strong>for</strong>m∞∑f(z) = a n (z − z 0 ) n . (24.59)n=−mNow consider the integral I of f(z) around a closed contour C that enclosesz = z 0 , but no other singular points. Using Cauchy’s theorem, this integral hasthe same value as the integral around a circle γ of radius ρ centred on z = z 0 ,since f(z) is analytic in the region between C <strong>and</strong> γ. On the circle we havez = z 0 + ρ exp iθ (<strong>and</strong> dz = iρ exp iθ dθ), <strong>and</strong> so∮I = f(z) dz==γ∞∑∮a n (z − z 0 ) n dzn=−m∞∑∫ 2πa n iρ n+1 exp[i(n +1)θ] dθ.n=−mFor every term in the series with n ≠ −1, we have∫ 2π[ iρiρ n+1 n+1 exp[i(n +1)θ]exp[i(n +1)θ] dθ =0i(n +1)but <strong>for</strong> the n = −1 termweobtain0∫ 2π0idθ=2πi.] 2π0=0,There<strong>for</strong>e only the term in (z − z 0 ) −1 contributes to the value of the integralaround γ (<strong>and</strong> there<strong>for</strong>e C), <strong>and</strong> I takes the value∮I = f(z) dz =2πia −1 . (24.60)CThus the integral around any closed contour containing a single pole of generalorder m (or, by extension, an essential singularity) is equal to 2πi times the residueof f(z) atz = z 0 .If we extend the above argument to the case where f(z) is continuous within<strong>and</strong> on a closed contour C <strong>and</strong> analytic, except <strong>for</strong> a finite number of poles,within C, then we arrive at the residue theorem∮f(z) dz =2πi ∑ R j , (24.61)Cjwhere ∑ j R j is the sum of the residues of f(z) at its poles within C.The method of proof is indicated by figure 24.13, in which (a) shows the originalcontour C referred to in (24.61) <strong>and</strong> (b) shows a contour C ′ giving the same value859


COMPLEX VARIABLESCC ′C(a)(b)Figure 24.13 The contours used to prove the residue theorem: (a) the originalcontour; (b) the contracted contour encircling each of the poles.to the integral, because f is analytic between C <strong>and</strong> C ′ . Now the contribution tothe C ′ integral from the polygon (a triangle <strong>for</strong> the case illustrated) joining thesmall circles is zero, since f is also analytic inside C ′ . Hence the whole value ofthe integral comes from the circles <strong>and</strong>, by result (24.60), each of these contributes2πi times the residue at the pole it encloses. All the circles are traversed in theirpositive sense if C is thus traversed <strong>and</strong> so the residue theorem follows. Formally,Cauchy’s theorem (24.40) is a particular case of (24.61) in which C encloses nopoles.Finally we prove another important result, <strong>for</strong> later use. Suppose that f(z) hasa simple pole at z = z 0 <strong>and</strong> so may be exp<strong>and</strong>ed as the Laurent seriesf(z) =φ(z)+a −1 (z − z 0 ) −1 ,where φ(z) is analytic within some neighbourhood surrounding z 0 . We wish tofind an expression <strong>for</strong> the integral I of f(z) along an open contour C, whichisthe arc of a circle of radius ρ centred on z = z 0 given by|z − z 0 | = ρ, θ 1 ≤ arg(z − z 0 ) ≤ θ 2 , (24.62)where ρ is chosen small enough that no singularity of f, other than z = z 0 , lieswithin the circle. Then I is given by∫∫∫I = f(z) dz = φ(z) dz + a −1 (z − z 0 ) −1 dz.CCIf the radius of the arc C is now allowed to tend to zero, then the first integraltends to zero, since the path becomes of zero length <strong>and</strong> φ is analytic <strong>and</strong>there<strong>for</strong>e continuous along it. On C, z = ρe iθ <strong>and</strong> hence the required expression<strong>for</strong> I is( ∫ θ2)1I = lim f(z) dz = lim a −1 ρ→0∫Cρ→0 θ 1ρe iθ iρeiθ dθ = ia −1 (θ 2 − θ 1 ). (24.63)860C


24.13 DEFINITE INTEGRALS USING CONTOUR INTEGRATIONWe note that result (24.60) is a special case of (24.63) in which θ 2 is equal toθ 1 +2π.24.13 Definite integrals using contour integrationThe remainder of this chapter is devoted to methods of applying contour integration<strong>and</strong> the residue theorem to various types of definite integral. However, threeapplications of contour integration, in which obtaining a value <strong>for</strong> the integral isnot the prime purpose of the exercise, have been postponed until chapter 25. Theyare the location of the zeros of a complex polynomial, the evaluation of the sumsof certain infinite series <strong>and</strong> the determination of inverse Laplace trans<strong>for</strong>ms.For the integral evalations considered here, not much preamble is given since,<strong>for</strong> this material, the simplest explanation is felt to be via a series of workedexamples that can be used as models.Suppose that an integral of the <strong>for</strong>m24.13.1 Integrals of sinusoidal functions∫ 2π0F(cos θ, sin θ) dθ (24.64)is to be evaluated. It can be made into a contour integral around the unit circleC by writing z =expiθ, <strong>and</strong> hencecos θ = 1 2 (z + z−1 ), sin θ = − 1 2 i(z − z−1 ), dθ = −iz −1 dz. (24.65)This contour integral can then be evaluated using the residue theorem, providedthe trans<strong>for</strong>med integr<strong>and</strong> has only a finite number of poles inside the unit circle<strong>and</strong>noneonit.◮Evaluate∫ 2πcos 2θI =dθ, b > a > 0. (24.66)0 a 2 + b 2 − 2ab cos θBy de Moivre’s theorem (section 3.4),cos nθ = 1 2 (zn + z −n ). (24.67)Using n = 2 in (24.67) <strong>and</strong> straight<strong>for</strong>ward substitution <strong>for</strong> the other functions of θ in(24.66) givesI =i ∮z 4 +12ab C z 2 (z − a/b)(z − b/a) dz.Thus there are two poles inside C, a double pole at z = 0 <strong>and</strong> a simple pole at z = a/b(recall that b>a).We could find the residue of the integr<strong>and</strong> at z = 0 by exp<strong>and</strong>ing the integr<strong>and</strong> as aLaurent series in z <strong>and</strong> identifying the coefficient of z −1 . Alternatively, we may use the861


COMPLEX VARIABLES<strong>for</strong>mula (24.56) with m = 2. Choosing the latter method <strong>and</strong> denoting the integr<strong>and</strong> byf(z), we haveddz [z2 f(z)] = d []z 4 +1dz (z − a/b)(z − b/a)= (z − a/b)(z − b/a)4z3 − (z 4 +1)[(z − a/b)+(z − b/a)].(z − a/b) 2 (z − b/a) 2Now setting z = 0 <strong>and</strong> applying (24.56), we findR(0) = a b + b a .For the simple pole at z = a/b, equation (24.57) gives the residue as[ ] (a/b) 4 +1R(a/b) = lim (z − a/b)f(z) =z→(a/b)(a/b) 2 (a/b − b/a)= − a4 + b 4ab(b 2 − a 2 ) .There<strong>for</strong>e by the residue theoremI =2πi ×i [ a 2 + b 22ab ab]− a4 + b 4=ab(b 2 − a 2 )2πa 2b 2 (b 2 − a 2 ) . ◭24.13.2 Some infinite integralsWe next consider the evaluation of an integral of the <strong>for</strong>m∫ ∞−∞where f(z) has the following properties:f(x) dx,(i) f(z) is analytic in the upper half-plane, Im z ≥ 0, except <strong>for</strong> a finite numberof poles, none of which is on the real axis;(ii) on a semicircle Γ of radius R (figure 24.14), R times the maximum of |f|on Γ tends to zero as R →∞(a sufficient condition is that zf(z) → 0as|z| →∞);(iii) ∫ 0−∞ f(x) dx <strong>and</strong> ∫ ∞0f(x) dx both exist.Since∫∣ f(z) dz∣ ≤ 2πR × (maximum of |f| on Γ),Γcondition (ii) ensures that the integral along Γ tends to zero as R →∞, afterwhich it is obvious from the residue theorem that the required integral is givenby∫ ∞−∞f(x) dx =2πi × (sum of the residues at poles with Im z>0).(24.68)862


24.13 DEFINITE INTEGRALS USING CONTOUR INTEGRATIONyΓ−RORxFigure 24.14A semicircular contour in the upper half-plane.◮Evaluate∫ ∞dxI =, where a is real.0 (x 2 + a 2 )4The complex function (z 2 + a 2 ) −4 has poles of order 4 at z = ±ai, ofwhichonlyz = aiis in the upper half-plane. Conditions (ii) <strong>and</strong> (iii) are clearly satisfied. For higher-orderpoles, <strong>for</strong>mula (24.56) <strong>for</strong> evaluating residues can be tiresome to apply. So, instead, we putz = ai + ξ <strong>and</strong> exp<strong>and</strong> <strong>for</strong> small ξ to obtain §The coefficient of ξ −1 is given by(1(z 2 + a 2 ) = 14 (2aiξ + ξ 2 ) = 11 − iξ ) −4.4 (2aiξ) 4 2a1 (−4)(−5)(−6)(2a) 4 3!( ) 3 −i= −5i2a 32a , 7<strong>and</strong> hence by the residue theorem∫ ∞dx−∞ (x 2 + a 2 ) = 10π4 32a , 7<strong>and</strong> so I =5π/(32a 7 ). ◭Condition (i) of the previous method required there to be no poles of theintegr<strong>and</strong> on the real axis, but in fact simple poles on the real axis can beaccommodated by indenting the contour as shown in figure 24.15. The indentationat the pole z = z 0 is in the <strong>for</strong>m of a semicircle γ of radius ρ in the upper halfplane,thus excluding the pole from the interior of the contour.§ This illustrates another useful technique <strong>for</strong> determining residues.863


COMPLEX VARIABLESyΓγ−RORxFigure 24.15 An indented contour used when the integr<strong>and</strong> has a simplepole on the real axis.What is then obtained from a contour integration, apart from the contributions<strong>for</strong> Γ <strong>and</strong> γ, is called the principal value of the integral, defined as ρ → 0byP∫ R−Rf(x) dx ≡∫ z0−ρ−Rf(x) dx +∫ Rz 0+ρf(x) dx.The remainder of the calculation goes through as be<strong>for</strong>e, but the contributionfrom the semicircle, γ, must be included. Result (24.63) of section 24.12 showsthat since only a simple pole is involved its contribution is−ia −1 π, (24.69)where a −1 is the residue at the pole <strong>and</strong> the minus sign arises because γ istraversed in the clockwise (negative) sense.We defer giving an example of an indented contour until we have establishedJordan’s lemma; we will then work through an example illustrating both. Jordan’slemma enables infinite integrals involving sinusoidal functions to be evaluated.For a function f(z) of a complex variable z, if(i) f(z) is analytic in the upper half-plane except <strong>for</strong> a finite number of poles in Im z>0,(ii) the maximum of |f(z)| →0 as |z| →∞in the upper half-plane,(iii) m>0,then∫I Γ = e imz f(z) dz → 0 as R →∞, (24.70)Γwhere Γ is the same semicircular contour as in figure 24.14.Note that this condition (ii) is less stringent than the earlier condition (ii) (see thestart of this section), since we now only require M(R) → 0 <strong>and</strong> not RM(R) → 0,where M is the maximum § of |f(z)| on |z| = R.§ More strictly, the least upper bound.864


24.13 DEFINITE INTEGRALS USING CONTOUR INTEGRATIONThe proof of the lemma is straight<strong>for</strong>ward once it has been observed that, <strong>for</strong>0 ≤ θ ≤ π/2,1 ≥ sin θ ≥ 2 θ π . (24.71)Then, since on Γ we have | exp(imz)| = | exp(−mR sin θ)|,∫∫ π∫ π/2I Γ ≤ |e imz f(z)| |dz| ≤MR e −mR sin θ dθ =2MR e −mR sin θ dθ.ΓThus, using (24.71),∫ π/2I Γ ≤ 2MR e −mR(2θ/π) dθ = πM (1 − e−mR ) < πM0mm ;hence, as R →∞, I Γ tends to zero since M tends to zero.◮Find the principal value of∫ ∞−∞cos mxx − a0dx, <strong>for</strong> a real, m>0.Consider the function (z − a) −1 exp(imz); although it has no poles in the upper half-planeit does have a simple pole at z = a, <strong>and</strong> further |(z − a) −1 |→0as|z| →∞. We will use acontour like that shown in figure 24.15 <strong>and</strong> apply the residue theorem. Symbolically,∫ a−ρ ∫ ∫ R ∫+ + + =0. (24.72)−RγNow as R →∞<strong>and</strong> ρ → 0 we have ∫ → 0, by Jordan’s lemma, <strong>and</strong> from (24.68) <strong>and</strong>Γ(24.69) we obtain∫ ∞e imxP−∞ x − a dx − iπa −1 =0, (24.73)where a −1 is the residue of (z − a) −1 exp(imz) atz = a, which is exp(ima). Then taking thereal <strong>and</strong> imaginary parts of (24.73) givesPP∫ ∞−∞∫ ∞−∞a+ρcos mxdx = −π sin ma,x − asin mxdx = π cos ma,x − aΓ0as required,as a bonus. ◭24.13.3 Integrals of multivalued functionsWe have discussed briefly some of the properties <strong>and</strong> difficulties associated withcertain multivalued functions such as z 1/2 or Ln z. It was mentioned that onemethod of managing such functions is by means of a ‘cut plane’. A similartechnique can be used with advantage to evaluate some kinds of infinite integralinvolving real functions <strong>for</strong> which the corresponding complex functions are multivalued.A typical contour employed <strong>for</strong> functions with a single branch point865


COMPLEX VARIABLESyΓγABCDxFigure 24.16 A typical cut-plane contour <strong>for</strong> use with multivalued functionsthat have a single branch point located at the origin.located at the origin is shown in figure 24.16. Here Γ is a large circle of radius R<strong>and</strong> γ is a small one of radius ρ, both centred on the origin. Eventually we willlet R →∞<strong>and</strong> ρ → 0.The success of the method is due to the fact that because the integr<strong>and</strong> ismultivalued, its values along the two lines AB <strong>and</strong> CD joining z = ρ to z = Rare not equal <strong>and</strong> opposite although both are related to the corresponding realintegral. Again an example provides the best explanation.◮Evaluate∫ ∞dxI =, a > 0.0 (x + a) 3 x1/2 We consider the integr<strong>and</strong> f(z) =(z + a) −3 z −1/2 <strong>and</strong> note that |zf(z)| →0onthetwocircles as ρ → 0<strong>and</strong>R →∞. Thus the two circles make no contribution to the contourintegral.The only pole of the integr<strong>and</strong> inside the contour is at z = −a (<strong>and</strong> is of order 3).To determine its residue we put z = −a + ξ <strong>and</strong> exp<strong>and</strong> (noting that (−a) 1/2 equalsa 1/2 exp(iπ/2) = ia 1/2 ):1(z + a) 3 z = 11/2 ξ 3 ia 1/2 (1 − ξ/a) 1/2= 1 (1+ 1 ξiξ 3 a 1/2 2 a + 3 ξ ···)28 a + .2The residue is thus −3i/(8a 5/2 ).The residue theorem (24.61) now gives∫ ∫ ∫+ +ABΓDC∫ ( ) −3i+ =2πi .γ 8a 5/2866


24.14 EXERCISESWe have seen that ∫ <strong>and</strong> ∫ vanish, <strong>and</strong> if we denote z by x along the line AB then itΓ γhas the value z = x exp 2πi along the line DC (note that exp 2πi must not be set equal to1 until after the substitution <strong>for</strong> z has been made in ∫ ). Substituting these expressions,DCThus<strong>and</strong>∫ ∞0∫dx0(x + a) 3 x + 1/2∞dx[x exp 2πi + a] 3 x 1/2 exp( 1 2(1 − 1 ) ∫ ∞dx 3π=exp πi 0 (x + a) 3 x1/2 4a 5/2I = 1 2 ×3π4a 5/2 . ◭2πi)=3π4a 5/2 .Several other examples of integrals of multivalued functions around a varietyof contours are included in the exercises that follow.24.14 Exercises24.1 Find an analytic function of z = x + iy whose imaginary part is(y cos y + x sin y)expx.24.2 Find a function f(z), analytic in a suitable part of the Arg<strong>and</strong> diagram, <strong>for</strong> whichsin 2xRe f =cosh 2y − cos 2x .Wherearethesingularitiesoff(z)?24.3 Find the radii of convergence of the following Taylor series:∞∑ z n(a)ln n , (b) ∑ ∞n!z nn , n n=2n=1∞∑∞∑( ) n 2n + p(c) z n n ln n , (d)z n , with p real.nn=1n=124.4 Find the Taylor series expansion about the origin of the function f(z) defined by∞∑ ( pz)f(z) = (−1) r+1 sin ,rr=1where p is a constant. Hence verify that f(z) is a convergent series <strong>for</strong> all z.24.5 Determine the types of singularities (if any) possessed by the following functionsat z =0<strong>and</strong>z = ∞:(a) (z − 2) −1 , (b) (1 + z 3 )/z 2 , (c) sinh(1/z),(d) e z /z 3 , (e) z 1/2 /(1 + z 2 ) 1/2 .24.6 Identify the zeros, poles <strong>and</strong> essential singularities of the following functions:(a) tan z, (b) [(z − 2)/z 2 ] sin[1/(1 − z)], (c) exp(1/z),(d) tan(1/z), (e) z 2/3 .867


COMPLEX VARIABLES24.7 Find the real <strong>and</strong> imaginary parts of the functions (i) z 2 , (ii) e z , <strong>and</strong> (iii) cosh πz.By considering the values taken by these parts on the boundaries of the region0 ≤ x, y ≤ 1, determine the solution of Laplace’s equation in that region thatsatisfies the boundary conditionsφ(x, 0) = 0, φ(0,y)=0,φ(x, 1) = x, φ(1,y)=y +sinπy.24.8 Show that the trans<strong>for</strong>mation∫ z1w =dζ0 (ζ 3 − ζ)1/2trans<strong>for</strong>ms the upper half-plane into the interior of a square that has one cornerat the origin of the w-plane <strong>and</strong> sides of length L, whereL =∫ π/20cosec 1/2 θdθ.24.9 The fundamental theorem of algebra states that, <strong>for</strong> a complex polynomial p n (z)of degree n, the equation p n (z) = 0 has precisely n complex roots. By applyingLiouville’s theorem (see the end of section 24.10) to f(z) =1/p n (z), prove thatp n (z) = 0 has at least one complex root. Factor out that root to obtain p n−1 (z)<strong>and</strong>, by repeating the process, prove the above theorem.24.10 Show that, if a is a positive real constant, the function exp(iaz 2 ) is analytic <strong>and</strong>→ 0as|z| →∞<strong>for</strong> 0 < arg z ≤ π/4. By applying Cauchy’s theorem to a suitablecontour prove that24.11 The function∫ ∞0cos(ax 2 ) dx =√ π8a .f(z) =(1− z 2 ) 1/2of the complex variable z is defined to be real <strong>and</strong> positive on the real axis inthe range −1


24.14 EXERCISES24.14 Prove that, <strong>for</strong> α>0, the integral∫ ∞t sin αt0 1+t dt 2has the value (π/2) exp(−α).24.15 Prove that∫ ∞cos mx0 4x 4 +5x 2 +1 dx = π (4e −m/2 − e −m) <strong>for</strong> m>0.624.16 Show that the principal value of the integral∫ ∞cos(x/a)dx−∞ x 2 − a 2is −(π/a)sin1.24.17 The following is an alternative (<strong>and</strong> roundabout!) way of evaluating the Gaussianintegral.(a) Prove that the integral of [exp(iπz 2 )]cosec πz around the parallelogram withcorners ±1/2 ± R exp(iπ/4) has the value 2i.(b) Show that the parts of the contour parallel to the real axis do not contributewhen R →∞.(c) Evaluate the integrals along the other two sides by putting z ′ = r exp(iπ/4)<strong>and</strong> working in terms of z ′ + 1 <strong>and</strong> 2z′ − 1 . Hence, by letting R →∞show2that∫ ∞−∞e −πr2 dr =1.24.18 By applying the residue theorem around a wedge-shaped contour of angle 2π/n,with one side along the real axis, prove that the integral∫ ∞dx0 1+x , nwhere n is real <strong>and</strong> ≥ 2, has the value (π/n)cosec (π/n).24.19 Using a suitable cut plane, prove that if α is real <strong>and</strong> 0


COMPLEX VARIABLES24.22 The equation of an ellipse in plane polar coordinates r, θ, with one of its foci atthe origin, isl=1− ɛ cos θ,rwhere l is a length (that of the latus rectum) <strong>and</strong> ɛ (0


25Applications of complex variablesIn chapter 24, we developed the basic theory of the functions of a complexvariable, z = x + iy, studied their analyticity (differentiability) properties <strong>and</strong>derived a number of results concerned with values of contour integrals in thecomplex plane. In this current chapter we will show how some of those results<strong>and</strong> properties can be exploited to tackle problems arising directly from physicalsituations or from apparently unrelated parts of mathematics.In the <strong>for</strong>mer category will be the use of the differential properties of the real<strong>and</strong> imaginary parts of a function of a complex variable to solve problems involvingLaplace’s equation in two dimensions, whilst an example of the latter mightbe the summation of certain types of infinite series. Other applications, such asthe Bromwich inversion <strong>for</strong>mula <strong>for</strong> Laplace trans<strong>for</strong>ms, appear as mathematicalproblems that have their origins in physical applications; the Bromwich inversionenables us to extract the spatial or temporal response of a system to an initialinput from the representation of that response in ‘frequency space’ – or, morecorrectly, imaginary frequency space.Other topics that will be considered are the location of the (complex) zeros ofa polynomial, the approximate evaluation of certain types of contour integralsusing the methods of steepest descent <strong>and</strong> stationary phase, <strong>and</strong> the so-called‘phase-integral’ solutions to some differential equations. For each of these a briefintroduction is given at the start of the relevant section <strong>and</strong> to repeat them herewould be pointless. We will there<strong>for</strong>e move on to our first topic of complexpotentials.25.1 Complex potentialsTowards the end of section 24.2 of the previous chapter it was shown that the real<strong>and</strong> the imaginary parts of an analytic function of z are separately solutions ofLaplace’s equation in two dimensions. Analytic functions thus offer a possible way871


APPLICATIONS OF COMPLEX VARIABLESyxFigure 25.1 The equipotentials (dashed circles) <strong>and</strong> field lines (solid lines)<strong>for</strong> a line charge perpendicular to the z-plane.of solving some two-dimensional physical problems describable by a potentialsatisfying ∇ 2 φ = 0. The general method is known as that of complex potentials.We also found that if f = u + iv is an analytic function of z then any curveu = constant intersects any curve v = constant at right angles. In the context ofsolutions of Laplace’s equation, this result implies that the real <strong>and</strong> imaginaryparts of f(z) have an additional connection between them, <strong>for</strong> if the set ofcontours on which one of them is a constant represents the equipotentials of asystem then the contours on which the other is constant, being orthogonal toeach of the first set, must represent the corresponding field lines or stream lines,depending on the context. The analytic function f is the complex potential. Itis conventional to use φ <strong>and</strong> ψ (rather than u <strong>and</strong> v) to denote the real <strong>and</strong>imaginary parts of a complex potential, so that f = φ + iψ.As an example, consider the functionf(z) = −q ln z (25.1)2πɛ 0in connection with the physical situation of a line charge of strength q per unitlength passing through the origin, perpendicular to the z-plane (figure 25.1). Itsreal <strong>and</strong> imaginary parts areφ = −q ln |z|, ψ = −q arg z. (25.2)2πɛ 0 2πɛ 0The contours in the z-plane of φ = constant are concentric circles <strong>and</strong> those ofψ = constant are radial lines. As expected these are orthogonal sets, but in additionthey are, respectively, the equipotentials <strong>and</strong> electric field lines appropriate to872


25.1 COMPLEX POTENTIALSthe field produced by the line charge. The minus sign is needed in (25.1) becausethe value of φ must decrease with increasing distance from the origin.Suppose we make the choice that the real part φ of the analytic function fgives the conventional potential function; ψ could equally well be selected. Thenwe may consider how the direction <strong>and</strong> magnitude of the field are related to f.◮Show that <strong>for</strong> any complex (electrostatic) potential f(z) the strength of the electric fieldis given by E = |f ′ (z)| <strong>and</strong> that its direction makes an angle of π − arg[f ′ (z)] with thex-axis.Because φ = constant is an equipotential, the field has componentsE x = − ∂φ∂x<strong>and</strong>E y = − ∂φ∂y . (25.3)Since f is analytic, (i) we may use the Cauchy–Riemann relations (24.5) to change thesecond of these, obtainingE x = − ∂φ∂x<strong>and</strong>E y = ∂ψ∂x ; (25.4)(ii) the direction of differentiation at a point is immaterial <strong>and</strong> sodfdz = ∂f∂x = ∂φ∂x + i ∂ψ∂x = −E x + iE y . (25.5)From these it can be seen that the field at a point is given in magnitude by E = |f ′ (z)|<strong>and</strong> that it makes an angle with the x-axis given by π − arg[f ′ (z)]. ◭It will be apparent from the above that much of physical interest can becalculated by working directly in terms of f <strong>and</strong> z. In particular, the electric fieldvector E may be represented, using (25.5) above, by the quantityE = E x + iE y = −[f ′ (z)] ∗ .Complex potentials can be used in two-dimensional fluid mechanics problemsin a similar way. If the flow is stationary (i.e. the velocity of the fluid does notdepend on time) <strong>and</strong> irrotational, <strong>and</strong> the fluid is both incompressible <strong>and</strong> nonviscous,then the velocity of the fluid can be described by V = ∇φ, whereφ is thevelocity potential <strong>and</strong> satisfies ∇ 2 φ = 0. If, <strong>for</strong> a complex potential f = φ + iψ,the real part φ is taken to represent the velocity potential then the curves ψ =constant will be the streamlines of the flow. In a direct parallel with the electricfield, the velocity may be represented in terms of the complex potential byV = V x + iV y =[f ′ (z)] ∗ ,the difference of a minus sign reflecting the same difference between the definitionsof E <strong>and</strong> V. The speed of the flow is equal to |f ′ (z)|. Points where f ′ (z) = 0, <strong>and</strong>thus the velocity is zero, are called stagnation points of the flow.Analogously to the electrostatic case, a line source of fluid at z = z 0 , perpendicularto the z-plane (i.e. a point from which fluid is emerging at a constant rate),873


APPLICATIONS OF COMPLEX VARIABLESis described by the complex potentialf(z) =k ln(z − z 0 ),where k is the strength of the source. A sink is similarly represented, but with kreplaced by −k. Other simple examples are as follows.(i) The flow of a fluid at a constant speed V 0 <strong>and</strong> at an angle α to the x-axisis described by f(z) =V 0 (exp iα)z.(ii) Vortex flow, in which fluid flows azimuthally in an anticlockwise directionaround some point z 0 , the speed of the flow being inversely proportionalto the distance from z 0 , is described by f(z) =−ik ln(z − z 0 ), where k isthe strength of the vortex. For a clockwise vortex k is replaced by −k.◮ Verify that the complex potential( )f(z) =V 0 z + a2zis appropriate to a circular cylinder of radius a placed so that it is perpendicular to auni<strong>for</strong>m fluid flow of speed V 0 parallel to the x-axis.Firstly, since f(z) isanalyticexceptatz = 0, both its real <strong>and</strong> imaginary parts satisfyLaplace’s equation in the region exterior to the cylinder. Also f(z) → V 0 z as z →∞,sothat Re f(z) → V 0 x, which is appropriate to a uni<strong>for</strong>m flow of speed V 0 in the x-directionfar from the cylinder.Writing z = r exp iθ <strong>and</strong> using de Moivre’s theorem we have[]f(z) =V 0 r exp iθ + a2r exp(−iθ) ( ) )= V 0 r + a2cos θ + iV 0(r − a2sin θ.rrThus we see that the streamlines of the flow described by f(z) are given by( )ψ = V 0 r − a2sin θ =constant.rIn particular, ψ =0onr = a, independently of the value of θ, <strong>and</strong>sor = a must be astreamline. Since there can be no flow of fluid across streamlines, r = a must correspondto a boundary along which the fluid flows tangentially. Thus f(z) is a solution of Laplace’sequation that satisfies all the physical boundary conditions of the problem, <strong>and</strong> so, by theuniqueness theorem, it is the appropriate complex potential. ◭By a similar argument, the complex potential f(z) =−E(z − a 2 /z) (note theminus signs) is appropriate to a conducting circular cylinder of radius a placedperpendicular to a uni<strong>for</strong>m electric field E in the x-direction.The real <strong>and</strong> imaginary parts of a complex potential f = φ + iψ have anotherinteresting relationship in the context of Laplace’s equation in electrostatics orfluid mechanics. Let us choose φ as the conventional potential, so that ψ representsthe stream function (or electric field, depending on the application), <strong>and</strong> consider874


25.1 COMPLEX POTENTIALSyQxPˆnFigure 25.2 A curve joining the points P <strong>and</strong> Q. Also shown is ˆn, the unitvector normal to the curve.the difference in the values of ψ at any two points P <strong>and</strong> Q connected by somepath C, as shown in figure 25.2. This difference is given by∫ Q ∫ Q( )∂ψ ∂ψψ(Q) − ψ(P )= dψ = dx +PP ∂x ∂y dy ,which, on using the Cauchy–Riemann relations, becomes∫ Q(ψ(Q) − ψ(P )= − ∂φ)∂φdx +P ∂y ∂x dy∫ Q∫ Q∂φ= ∇φ · ˆn ds =PP ∂n ds,where ˆn is the vector unit normal to the path C <strong>and</strong> s is the arc length along thepath; the last equality is written in terms of the normal derivative ∂φ/∂n ≡∇φ· ˆn.Now suppose that in an electrostatics application, the path C is the surface ofa conductor; then∂φ∂n = − σ ,ɛ 0where σ is the surface charge density per unit length normal to the xy-plane.There<strong>for</strong>e −ɛ 0 [ψ(Q) − ψ(P )] is equal to the charge per unit length normal to thexy-plane on the surface of the conductor between the points P <strong>and</strong> Q. Similarly,in fluid mechanics applications, if the density of the fluid is ρ <strong>and</strong> its velocity isV then∫ Q∫ Qρ[ψ(Q) − ψ(P )] = ρ ∇φ · ˆn ds = ρ V · ˆn dsPis equal to the mass flux between P <strong>and</strong> Q per unit length perpendicular to thexy-plane.875P


APPLICATIONS OF COMPLEX VARIABLES◮ A conducting circular cylinder of radius a is placed with its centre line passing throughthe origin <strong>and</strong> perpendicular to a uni<strong>for</strong>m electric field E in the x-direction. Find the chargeper unit length induced on the half of the cylinder that lies in the region x


25.2 APPLICATIONS OF CONFORMAL TRANSFORMATIONSyssxrr(a) z-plane (b) w-plane (c) w-planeFigure 25.3 The equipotential lines (broken) <strong>and</strong> field lines (solid) (a) <strong>for</strong> aninfinite charged conducting plane at y =0,wherez = x + iy, <strong>and</strong> after thetrans<strong>for</strong>mations (b) w = z 2 <strong>and</strong> (c) w = z 1/2 of the situation shown in (a).◮ Find the complex electrostatic potential associated with an infinite charged conductingplate y =0, <strong>and</strong> thus obtain those associated with(i) a semi-infinite charged conducting plate (r >0, s =0);(ii) the inside of a right-angled charged conducting wedge (r > 0, s = 0 <strong>and</strong>r =0, s>0).Figure 25.3(a) shows the equipotentials (broken lines) <strong>and</strong> field lines (solid lines) <strong>for</strong> theinfinite charged conducting plane y = 0. Suppose that we elect to make the real part ofthe complex potential coincide with the conventional electrostatic potential. If the plate ischarged to a potential V then clearlyφ(x, y) =V − ky, (25.6)where k is related to the charge density σ by k = σ/ɛ 0 , since physically the electric field Ehas components (0,σ/ɛ 0 )<strong>and</strong>E = −∇φ.Thus what is needed is an analytic function of z, ofwhichtherealpartisV − ky. Thiscan be obtained by inspection, but we may proceed <strong>for</strong>mally <strong>and</strong> use the Cauchy–Riemannrelations to obtain the imaginary part ψ(x, y) as follows:∂ψ∂y = ∂φ∂x =0 <strong>and</strong> ∂ψ∂x = − ∂φ∂y = k.Hence ψ = kx + c <strong>and</strong>, absorbing c into V , the required complex potential isf(z) =V − ky + ikx = V + ikz. (25.7)(i) Now consider the trans<strong>for</strong>mationw = g(z) =z 2 . (25.8)This satisfies the criteria <strong>for</strong> a con<strong>for</strong>mal mapping (except at z = 0) <strong>and</strong> carries the upperhalf of the z-plane into the entire w-plane; the equipotential plane y =0goesintothehalf-plane r>0, s =0.By the general results proved, f(z), when expressed in terms of r <strong>and</strong> s, will give acomplex potential whose real part will be constant on the half-plane in question; we877


APPLICATIONS OF COMPLEX VARIABLESdeduce thatF(w) =f(z) =V + ikz = V + ikw 1/2 (25.9)is the required potential. Expressed in terms of r, s <strong>and</strong> ρ =(r 2 + s 2 ) 1/2 , w 1/2 is given by[ ( ) 1/2 ( ) ] 1/2 ρ + r ρ − rw 1/2 = ρ 1/2 + i, (25.10)2ρ2ρ<strong>and</strong>, in particular, the electrostatic potential is given byΦ(r, s) =ReF(w) =V −k √2[(r 2 + s 2 ) 1/2 − r ] 1/2. (25.11)The corresponding equipotentials <strong>and</strong> field lines are shown in figure 25.3(b). Using results(25.3)–(25.5), the magnitude of the electric field is|E| = |F ′ (w)| = | 1 2 ikw−1/2 | = 1 2 k(r2 + s 2 ) −1/4 .(ii) A trans<strong>for</strong>mation ‘converse’ to that used in (i),w = g(z) =z 1/2 ,has the effect of mapping the upper half of the z-plane into the first quadrant of thew-plane <strong>and</strong> the conducting plane y =0intothewedger>0, s =0<strong>and</strong>r =0,s>0.The complex potential now becomesF(w) =V + ikw 2= V + ik[(r 2 − s 2 )+2irs], (25.12)showing that the electrostatic potential is V −2krs <strong>and</strong> that the electric field has componentsE =(2ks,2kr). (25.13)Figure 25.3(c) indicates the approximate equipotentials <strong>and</strong> field lines. (Note that, in bothtrans<strong>for</strong>mations, g ′ (z) iseither0or∞ at the origin, <strong>and</strong> so neither trans<strong>for</strong>mation iscon<strong>for</strong>mal there. Consequently there is no violation of result (ii), given at the start ofsection 24.7, concerning the angles between intersecting lines.) ◭The method of images, discussed in section 21.5, can be used in conjunction withcon<strong>for</strong>mal trans<strong>for</strong>mations to solve some problems involving Laplace’s equationin two dimensions.◮A wedge of angle π/α with its vertex at z =0is <strong>for</strong>med by two semi-infinite conductingplates, as shown in figure 25.4(a). A line charge of strength q per unit length is positionedat z = z 0 , perpendicular to the z-plane. By considering the trans<strong>for</strong>mation w = z α ,findthecomplex electrostatic potential <strong>for</strong> this situation.Let us consider the action of the trans<strong>for</strong>mation w = z α on the lines defining the positionsof the conducting plates. The plate that lies along the positive x-axis is mapped onto thepositive r-axis in the w-plane, whereas the plate that lies along the direction exp(iπ/α) ismapped into the negative r-axis, as shown in figure 25.4(b). Similarly the line charge at z 0is mapped onto the point w 0 = z0 α.From figure 25.4(b), we see that in the w-plane the problem can be solved by introducinga second line charge of opposite sign at the point w0 ∗ , so that the potential Φ = 0 alongthe r-axis. The complex potential <strong>for</strong> such an arrangement is simplyF(w) =−q ln(w − w 0 )+q ln(w − w2πɛ 0 2πɛ0).∗0878


25.3 LOCATION OF ZEROSφ =0yπ/αz 0w = z α sw 0φ =0xΦ=0Φ=0r(a)(b)w ∗ 0Figure 25.4 (a) An infinite conducting wedge with interior angle π/α <strong>and</strong> aline charge at z = z 0 ; (b) after the trans<strong>for</strong>mation w = z α , with an additionalimagechargeplacedatw = w0 ∗.Substituting w = z α into the above shows that the required complex potential in theoriginal z-plane isf(z) =q ( ) z α − z0∗αln. ◭2πɛ 0 z α − z0αIt should be noted that the appearance of a complex conjugate in the finalexpression is not in conflict with the general requirement that the complexpotential be analytic. It is z ∗ that must not appear; here, z0 ∗α is no more than aparameter of the problem.25.3 Location of zerosThe residue theorem, relating the value of a closed contour integral to the sum ofthe residues at the poles enclosed by the contour, was discussed in the previouschapter. One important practical use of an extension to the theorem is that oflocating the zeros of functions of a complex variable. The location of such zeroshas a particular application in electrical network <strong>and</strong> general oscillation theory,since the complex zeros of certain functions (usually polynomials) give the systemparameters (usually frequencies) at which system instabilities occur. As the basisof a method <strong>for</strong> locating these zeros we next prove three important theorems.(i) If f(z) has poles as its only singularities inside a closed contour C <strong>and</strong> isnot zero at any point on C then∮f ′ (z)C f(z) dz =2πi ∑ (N j − P j ). (25.14)jHere N j is the order of the jth zero of f(z) enclosed by C. Similarly P j is theorder of the jth pole of f(z) inside C.To prove this we note that, at each position z j , f(z) can be written asf(z) =(z − z j ) mj φ(z), (25.15)879


APPLICATIONS OF COMPLEX VARIABLESwhere φ(z) is analytic <strong>and</strong> non-zero at z = z j <strong>and</strong> m j is positive <strong>for</strong> a zero <strong>and</strong>negative <strong>for</strong> a pole. Then the integr<strong>and</strong> f ′ (z)/f(z) takes the <strong>for</strong>mf ′ (z)f(z) = m j+ φ′ (z)z − z j φ(z) . (25.16)Since φ(z j ) ≠ 0, the second term on the RHS is analytic; thus the integr<strong>and</strong> has asimple pole at z = z j , with residue m j . For zeros m j = N j <strong>and</strong> <strong>for</strong> poles m j = −P j ,<strong>and</strong> thus (25.14) follows from the residue theorem.(ii) If f(z) is analytic inside C <strong>and</strong> not zero at any point on it then2π ∑ jN j =∆ C [arg f(z)], (25.17)where ∆ C [x] denotes the variation in x around the contour C.Since f is analytic, there are no P j ; further, sincef ′ (z)f(z) = d [Ln f(z)], (25.18)dzequation (25.14) can be writtenHowever,2πi ∑ ∮N j =Cf ′ (z)f(z) dz =∆ C[Ln f(z)]. (25.19)∆ C [Ln f(z)] = ∆ C [ln |f(z)|]+i∆ C [arg f(z)], (25.20)<strong>and</strong>, since C is a closed contour, ln |f(z)| must return to its original value; so thereal term on the RHS is zero. Comparison of (25.19) <strong>and</strong> (25.20) then establishes(25.17), which is known as the principle of the argument.(iii) If f(z) <strong>and</strong> g(z) are analytic within <strong>and</strong> on a closed contour C <strong>and</strong>|g(z)| < |f(z)| on C then f(z) <strong>and</strong>f(z) +g(z) have the same number of zerosinside C; thisisRouché’s theorem.With the conditions given, neither f(z) norf(z)+g(z) can have a zero on C.So, applying theorem (ii) with an obvious notation,2π ∑ j N j(f + g) =∆ C [arg(f + g)]=∆ C [arg f]+∆ C [arg(1 + g/f)]=2π ∑ k N k(f)+∆ C [arg(1 + g/f)]. (25.21)Further, since |g| < |f| on C, 1+g/f always lies within a unit circle centredon z = 1; thus its argument always lies in the range −π/2 < arg(1 + g/f)


25.3 LOCATION OF ZEROSpolynomials, only the behaviour of a single term in the function need be consideredif the contour is chosen appropriately. For example, <strong>for</strong> a polynomial,f(z)+g(z) = ∑ N0 b iz i , only the properties of its largest power, taken as f(z), needbe investigated if a circular contour is chosen with radius R sufficiently large that,on the contour, the magnitude of the largest power term, |b N R N |, is greater thanthe sum of the magnitudes of all other terms. It is obvious that f(z) =b N z N hasN zeros inside |z| = R (all at the origin); consequently, f + g also has N zerosinside the same circle.The corresponding situation, in which only the properties of the polynomial’ssmallest power, again taken as f(z), need be investigated is a circular contourwith a radius R chosen sufficiently small that, on the contour, the magnitude ofthe smallest power term (usually the constant term in a polynomial) is greaterthan the sum of the magnitudes of all other terms. Then, a similar argument tothat given above shows that, since f(z) =b 0 has no zeros inside |z| = R, neitherdoes f + g.Aweak<strong>for</strong>mofthemaximum-modulus theorem may also be deduced. Thisstates that if f(z) is analytic within <strong>and</strong> on a simple closed contour C then |f(z)|attains its maximum value on the boundary of C. The proof is as follows.Let |f(z)| ≤M on C with equality at at least one point of C. Now supposethat there is a point z = a inside C such that |f(a)| >M. Then the functionh(z) ≡ f(a) is such that |h(z)| > |−f(z)| on C, <strong>and</strong> thus, by Rouché’s theorem,h(z) <strong>and</strong>h(z) − f(z) have the same number of zeros inside C. Buth(z) (≡ f(a))has no zeros inside C, <strong>and</strong>, again by Rouché’s theorem, this would imply thatf(a) − f(z) has no zeros in C. However, f(a) − f(z) clearly has a zero at z = a,<strong>and</strong> so we have a contradiction; the assumption of the existence of a point z = ainside C such that |f(a)| >Mmust be invalid. This establishes the theorem.The stronger <strong>for</strong>m of the maximum-modulus theorem, which we do not prove,states, in addition, that the maximum value of f(z) is not attained at any interiorpoint except <strong>for</strong> the case where f(z) is a constant.◮ Show that the four zeros of h(z) =z 4 + z +1 occur one in each quadrant of the Arg<strong>and</strong>diagram <strong>and</strong> that all four lie between the circles |z| =2/3 <strong>and</strong> |z| =3/2.Putting z = x <strong>and</strong> z = iy shows that no zeros occur on the real or imaginary axes. Theymust there<strong>for</strong>e occur in conjugate pairs, as can be shown by taking the complex conjugateof h(z) =0.Now take C as the contour OXY O shown in figure 25.5 <strong>and</strong> consider the changes∆[arg h] in the argument of h(z) asz traverses C.(i) OX: argh is everywhere zero, since h is real, <strong>and</strong> thus ∆ OX [arg h] =0.(ii) XY : z = R exp iθ <strong>and</strong>soargh changes by an amount∆ XY [arg h] =∆ XY [arg z 4 ]+∆ XY [arg(1 + z −3 + z −4 )]{=∆ XY [arg R 4 e 4iθ ]+∆ XY arg[1+O(R −3 )] }=2π +O(R −3 ). (25.22)881


APPLICATIONS OF COMPLEX VARIABLESyYROXxFigure 25.5 A contour <strong>for</strong> locating the zeros of a polynomial that occur inthe first quadrant of the Arg<strong>and</strong> diagram.(iii) YO: z = iy <strong>and</strong> so arg h =tan −1 y/(y 4 + 1), which starts at O(R −3 ) <strong>and</strong> finishes at0asy goes from large R to 0. It never reaches π/2 because y 4 +1=0hasnorealpositive root. Thus ∆ YO [arg h] =0.Hence <strong>for</strong> the complete contour ∆ C [arg h] =0+2π +0+O(R −3 ), <strong>and</strong>, if R is allowedto tend to infinity, we deduce from (25.17) that h(z) has one zero in the first quadrant.Furthermore, since the roots occur in conjugate pairs, a second root must lie in the fourthquadrant, <strong>and</strong> the other pair must lie in the second <strong>and</strong> third quadrants.To show that the zeros lie within the given annulus in the z-plane we apply Rouché’stheorem, as follows.(i) With C as |z| =3/2, f = z 4 , g = z +1.Now|f| =81/16 on C <strong>and</strong> |g| ≤1+|z|


25.4 SUMMATION OF SERIES◮By considering∮π cot πzC (a + z) dz, 2where a is not an integer <strong>and</strong> C is a circle of large radius, evaluate∞∑ 1(a + n) . 2n=−∞The integr<strong>and</strong> has (i) simple poles at z = integer n, <strong>for</strong>−∞


APPLICATIONS OF COMPLEX VARIABLESIm sLRe sλFigure 25.6 The integration path of the inverse Laplace trans<strong>for</strong>m is alongthe infinite line L. Thequantityλ must be positive <strong>and</strong> large enough <strong>for</strong> allpoles of the integr<strong>and</strong> to lie to the left of L.25.5 Inverse Laplace trans<strong>for</strong>mAs a further example of the use of contour integration we now discuss a methodwhereby the process of Laplace trans<strong>for</strong>mation, discussed in chapter 13, can beinverted.It will be recalled that the Laplace trans<strong>for</strong>m ¯f(s) ofafunctionf(x), x ≥ 0, isgiven by¯f(s) =∫ ∞0e −sx f(x) dx, Re s>s 0 . (25.24)In chapter 13, functions f(x) were deduced from the trans<strong>for</strong>ms by means of aprepared dictionary. However, an explicit <strong>for</strong>mula <strong>for</strong> an unknown inverse maybe written in the <strong>for</strong>m of an integral. It is known as the Bromwich integral <strong>and</strong> isgiven byf(x) = 12πi∫ λ+i∞λ−i∞e sx¯f(s) ds, λ > 0, (25.25)where s is treated as a complex variable <strong>and</strong> the integration is along the line Lindicated in figure 25.6. The position of the line is dictated by the requirementsthat λ is positive <strong>and</strong> that all singularities of ¯f(s) lie to the left of the line.That (25.25) really is the unique inverse of (25.24) is difficult to show <strong>for</strong> generalfunctions <strong>and</strong> trans<strong>for</strong>ms, but the following verification should at least make it884


25.5 INVERSE LAPLACE TRANSFORMΓRΓRΓRLLL(a) (b) (c)Figure 25.7 Some contour completions <strong>for</strong> the integration path L of theinverse Laplace trans<strong>for</strong>m. For details of when each is appropriate see themain text.plausible:∫ λ+i∞f(x) = 1∫ ∞ds e sx e −su f(u) du, Re(s) > 0, i.e. λ>0,2πi λ−i∞0= 1 ∫ ∞ ∫ λ+i∞du f(u) e s(x−u) ds2πi 0λ−i∞= 1 ∫ ∞ ∫ ∞du f(u) e λ(x−u) e ip(x−u) i dp, putting s = λ + ip,2πi 0−∞= 1 ∫ ∞f(u)e λ(x−u) 2πδ(x − u) du2π 0{ f(x) x ≥ 0,=(25.26)0 x


APPLICATIONS OF COMPLEX VARIABLESIdeally, we would like the contribution to the integral from the circular arc Γ totend to zero as its radius R →∞. Using a modified version of Jordan’s lemma,it may be shown that this is indeed the case if there exist constants M>0<strong>and</strong>α>0 such that on Γ|¯f(s)| ≤ M R α .Moreover, this condition always holds when ¯f(s) hasthe<strong>for</strong>m¯f(s) = P (s)Q(s) ,where P (s) <strong>and</strong>Q(s) are polynomials <strong>and</strong> the degree of Q(s) is greater than thatof P (s).When the contribution from the part-circle Γ tends to zero as R →∞,wehave from the residue theorem that the inverse Laplace trans<strong>for</strong>m (25.25) is givensimply byf(t) = ∑ (residues of ¯f(s)e sx at all poles ) . (25.27)◮Find the function f(x) whose Laplace trans<strong>for</strong>m iss¯f(s) =s 2 − k , 2where k is a constant.It is clear that ¯f(s) is of the <strong>for</strong>m required <strong>for</strong> the integral over the circular arc Γ to tendto zero as R →∞, <strong>and</strong> so we may use the result (25.27) directly. Now¯f(s)e sx se sx=(s − k)(s + k) ,<strong>and</strong> thus has simple poles at s = k <strong>and</strong> s = −k. Using (24.57) the residues at each pole canbe easily calculated asR(k) = kekx<strong>and</strong> R(−k) = ke−kx2k2k .Thus the inverse Laplace trans<strong>for</strong>m is given by(f(x) = 1 2 e kx + e −kx) =coshkx.This result may be checked by computing the <strong>for</strong>ward trans<strong>for</strong>m of cosh kx. ◭Sometimes a little more care is required when deciding in which half-plane toclose the contour C.◮Find the function f(x) whose Laplace trans<strong>for</strong>m is¯f(s) = 1 s (e−as − e −bs ),where a <strong>and</strong> b are fixed <strong>and</strong> positive, with b>a.From (25.25) we have the integralf(x) = 12πi∫ λ+i∞λ−i∞e (x−a)s − e (x−b)ssds. (25.28)886


25.5 INVERSE LAPLACE TRANSFORMf(x)1abxFigure 25.8b>a.The result of the Laplace inversion of ¯f(s) =s −1 (e −as −e −bs )withNow, despite appearances to the contrary, the integr<strong>and</strong> has no poles, as may be confirmedby exp<strong>and</strong>ing the exponentials as Taylor series about s = 0. Depending on the value of x,several cases arise.(i) For xb. (25.30)(iii) For a


APPLICATIONS OF COMPLEX VARIABLES25.6 Stokes’ equation <strong>and</strong> Airy integralsMuch of the analysis of situations occurring in physics <strong>and</strong> engineering is concernedwith what happens at a boundary within or surrounding a physical system.Sometimes the existence of a boundary imposes conditions on the behaviour ofvariables describing the state of the system; obvious examples include the zerodisplacement at its end-points of an anchored vibrating string <strong>and</strong> the zeropotential contour that must coincide with a grounded electrical conductor.More subtle are the effects at internal boundaries, where the same non-vanishingvariable has to describe the situation on either side of the boundary but itsbehaviour is quantitatively, or even qualitatively, different in the two regions. Inthis section we will study an equation, Stokes’ equation, whose solutions havethis latter property; as well as solutions written as series in the usual way, we willfind others expressed as complex integrals.The Stokes’ equation can be written in several <strong>for</strong>ms, e.g.d 2 yd+ λxy =0;dx2 2 yd+ xy =0;dx2 2 ydx 2 = xy.We will adopt the last of these, but write it asd 2 y= zy (25.32)dz2 to emphasis that its complex solutions are valid <strong>for</strong> a complex independentvariable z, though this also means that particular care has to be exercised whenexamining their behaviour in different parts of the complex z-plane. The other<strong>for</strong>ms of Stokes’ equation can all be reduced to that of (25.32) by suitable(complex) scaling of the independent variable.25.6.1 The solutions of Stokes’ equationIt will be immediately apparent that, even <strong>for</strong> z restricted to be real <strong>and</strong> denotedby x, the behaviour of the solutions to (25.32) will change markedly as x passesthrough x = 0. For positive x they will have similar characteristics to the solutionsof y ′′ = k 2 y,wherek is real; these have monotonic exponential <strong>for</strong>ms, eitherincreasing or decreasing. On the other h<strong>and</strong>, when x is negative the solutionswill be similar to those of y ′′ + k 2 y = 0, i.e. oscillatory functions of x. Thisisjust the sort of behaviour shown by the wavefunction describing light diffractedby a sharp edge or by the quantum wavefunction describing a particle near tothe boundary of a region which it is classically <strong>for</strong>bidden to enter on energygrounds. Other examples could be taken from the propagation of electromagneticradiation in an ion plasma or wave-guide.Let us examine in a bit more detail the behaviour of plots of possible solutionsy(z) of Stokes’ equation in the region near z = 0 <strong>and</strong>, in particular, what may888


25.6 STOKES’ EQUATION AND AIRY INTEGRALS(b)y(c)(a)(a)(c)z(b)Figure 25.9 Behaviour of the solutions y(z) ofStokes’equationnearz =0<strong>for</strong> various values of λ = −y ′ (0). (a) with λ small, (b) with λ large <strong>and</strong> (c) withλ appropriate to the Airy function Ai(z).happen in the region z>0. For definiteness <strong>and</strong> ease of illustration (see figure25.9), let us suppose that both y <strong>and</strong> z, <strong>and</strong> hence the derivatives of y, are real <strong>and</strong>that y(0) is positive; if it were negative, our conclusions would not be changedsince equation (25.32) is invariant under y(z) →−y(z). The only difference wouldbe that all plots of y(z) would be reflected in the z-axis.We first note that d 2 y/dx 2 , <strong>and</strong> hence also the curvature of the plot, has thesame sign as z, i.e. it has positive curvature when z>0, <strong>for</strong> so long as y(z)remains positive there. What will happen to the plot <strong>for</strong> z>0 there<strong>for</strong>e dependscrucially on the value of y ′ (0). If this slope is positive or only slightly negativethe positive curvature will carry the plot, either immediately or ultimately, furtheraway from the z-axis. On the other h<strong>and</strong>, if y ′ (0) is negative but sufficiently largein magnitude, the plot will cross the y = 0 line; if this happens the sign of thecurvature reverses <strong>and</strong> again the plot will be carried ever further from the z-axis,only this time towards large negative values.Between these two extremes it seems at least plausible that there is a particularnegative value of y ′ (0) that leads to a plot that approaches the z-axis asymptotically,never crosses it (<strong>and</strong> so always has positive curvature), <strong>and</strong> has a slopethat, whilst always negative, tends to zero in magnitude. There is such a solution,known as Ai(z), whose properties we will examine further in the followingsubsections. The three cases are illustrated in figure 25.9.The behaviour of the solutions of (25.32) in the region z


APPLICATIONS OF COMPLEX VARIABLES<strong>for</strong>ward, in that, whatever the sign of y at any particular point z, the curvaturealways has the opposite sign. Consequently the curve always bends towards thez-axis, crosses it, <strong>and</strong> then bends towards the axis again. Thus the curve exhibitsoscillatory behaviour. Furthermore, as −z increases, the curvature <strong>for</strong> any given|y| gets larger; as a consequence, the oscillations become increasingly more rapid<strong>and</strong> their amplitude decreases.25.6.2 Series solution of Stokes’ equationObtaining a series solution of Stokes’ equation presents no particular difficultywhen the methods of chapter 16 are used. The equation, written in the <strong>for</strong>md 2 y− zy =0,dz2 has no singular points except at z = ∞. Every other point in the z-plane isan ordinary point <strong>and</strong> so two linearly independent series expansions about it(<strong>for</strong>mally with indicial values σ = 0 <strong>and</strong> σ = 1) can be found. Those about z =0take the <strong>for</strong>ms ∑ ∞0 a nz n <strong>and</strong> ∑ ∞0 b nz n+1 . The corresponding recurrence relationsare(n +3)(n +2)a n+3 = a n <strong>and</strong> (n +4)(n +3)b n+3 = b n ,<strong>and</strong>thetwoseries(witha 0 = b 0 = 1) take the <strong>for</strong>msy 1 (z) =1+z3(3)(2) + z 6(6)(5)(3)(2) + ··· ,y 2 (z) =z +z4(4)(3) + z 7(7)(6)(4)(3) + ··· .The ratios of successive terms <strong>for</strong> the two series are thusa n+3 z n+3 z 3a n z n =(n +3)(n +2)<strong>and</strong>b n+3 z n+4 z 3b n z n+1 =(n +4)(n +3) .It follows from the ratio test that both series are absolutely convergent <strong>for</strong> all z.A similar argument shows that the series <strong>for</strong> their derivatives are also absolutelyconvergent <strong>for</strong> all z. Any solution of the Stokes’ equation is representable as asuperposition of the two series <strong>and</strong> so is analytic <strong>for</strong> all finite z; itisthere<strong>for</strong>eanintegral function with its only singularity at infinity.25.6.3 Contour integral solutionsWe now move on to another <strong>for</strong>m of solution of the Stokes’ equation (25.32), onethat takes the <strong>for</strong>m of a contour integral in which z appears as a parameter in890


25.6 STOKES’ EQUATION AND AIRY INTEGRALSthe integr<strong>and</strong>. Consider the contour integraly(z) =∫ baf(t) exp(zt) dt, (25.33)in which a, b <strong>and</strong> f(t) are all yet to be chosen. Note that the contour is in thecomplex t-plane <strong>and</strong> that the path from a to b can be distorted as required solong as no poles of the integr<strong>and</strong> are trapped between an original path <strong>and</strong> itsdistortion.Substitution of (25.33) into (25.32) yields∫ bat 2 f(t) exp(zt) dt =∫ bazf(t) exp(zt) dt= [ f(t) exp(zt) ] b a −df(t)exp(zt) dt.a dtIf we could choose the limits a <strong>and</strong> b so that the end-point contributions vanishthen Stokes’ equation would be satisfied by (25.33), provided f(t) satisfiesdf(t)+ t 2 f(t) =0dt⇒ f(t) =A exp(− 1 3 t3 ), (25.34)where A is any constant.To make the end-point contributions vanish we must choose a <strong>and</strong> b such thatexp(− 1 3 t3 + zt) = 0 <strong>for</strong> both values of t. This can only happen if |a| →∞<strong>and</strong>|b| →∞<strong>and</strong>, even then, only if Re (t 3 ) is positive. This condition is satisfied if2nπ − 1 2 π


APPLICATIONS OF COMPLEX VARIABLESIm tC 1C 2C 3Re tFigure 25.10 The contours used in the complex t-plane to define the functionsAi(z) <strong>and</strong>Bi(z).though sometimes the sense of traversal is reversed. Consequently there arerelationships connecting Ai <strong>and</strong> Bi when the rotated variables are used as theirarguments. As two examples,Ai(z)+ΩAi(Ωz)+Ω 2 Ai(Ω 2 z)=0, (25.37)Bi(z) =i[Ω 2 Ai(Ω 2 z) − ΩAi(Ωz)]=e −πi/6 Ai(ze −2πi/3 )+e πi/6 Ai(ze 2πi/3 ).(25.38)Since the only requirements <strong>for</strong> the integral paths is that they start <strong>and</strong> end inthe correct sectors, we can distort path C 1 so that it lies on the imaginary axis<strong>for</strong> virtually its whole length <strong>and</strong> just to the left of the axis at its two ends. Thisenables us to obtain an alternative expression <strong>for</strong> Ai(z), as follows.Setting t = is, wheres is real <strong>and</strong> −∞


25.6 STOKES’ EQUATION AND AIRY INTEGRALScontour integral (25.35) with the particular solution of Stokes’ equation thatdecays monotonically to zero <strong>for</strong> real z>0as|z| →∞. As discussed in subsection25.6.1, all solutions except the one called Ai(z) tendto±∞ as z (real) takes onincreasingly large positive values <strong>and</strong> so their asymptotic <strong>for</strong>ms reflect this. In aworked example in subsection 25.8.2 we use the method of steepest descents (asaddle-point method) to show that the function defined by (25.39) has exactlythe characteristic asymptotic property expected of Ai(z) (see page 911). It followsthat it is the same function as Ai(z), up to a real multiplicative constant.The choice of definition (25.36) as the other named solution Bi(z) ofStokes’equation is a less obvious one. However, it is made on the basis of its behaviour <strong>for</strong>negative real values of z. As discussed earlier, Ai(z) oscillates almost sinusoidallyin this region, except <strong>for</strong> a relatively slow increase in frequency <strong>and</strong> an evenslower decrease in amplitude as −z increases. The solution Bi(z) is chosen to bethe particular function that exhibits the same behaviour as Ai(z) exceptthatitisin quadrature with Ai, i.e. it is π/2 out of phase with it. Specifically, as x →−∞,Ai(x) ∼Bi(x) ∼(1 2|x|3/2√ sin + π ), (25.40)2πx1/4 3 4(1 2|x|3/2√ cos + π2πx1/4 3 4). (25.41)There is a close parallel between this choice <strong>and</strong> that of taking sine <strong>and</strong> cosinefunctions as the basic independent solutions of the simple harmonic oscillatorequation. Plots of Ai(z) <strong>and</strong>Bi(z) <strong>for</strong>realz are shown in figure 25.11.◮By choosing a suitable contour <strong>for</strong> C 1 in (25.35), express Ai(0) in terms of the gammafunction.With z set equal to zero, (25.35) takes the <strong>for</strong>mAi(0) = 1 ∫exp(− 1 32πit3 ) dt.C 1We again use the freedom to choose the specific line of the contour so as to make theactual integration as simple as possible.HereweconsiderC 1 as made up of two straight-line segments: one along the linearg t =4π/3, starting at infinity in the correct sector <strong>and</strong> ending at the origin; the otherstarting at the origin <strong>and</strong> going to infinity along the line arg t =2π/3, thus ending in thecorrect final sector. On each, we set 1 3 t3 = s, wheres is real <strong>and</strong> positive on both lines.Then dt = e 4πi/3 (3s) −2/3 ds on the first segment <strong>and</strong> dt = e 2πi/3 (3s) −2/3 ds on the second.893


APPLICATIONS OF COMPLEX VARIABLES10.5−10−5z−0.5Figure 25.11z.The functions Ai(z) (full line) <strong>and</strong> Bi(z) (broken line) <strong>for</strong> realThen we haveAi(0) = 12πi∫ 0∞∫ ∞= 3−2/3e −s e 4πi/3 (3s) −2/3 ds + 12πi∫ ∞e −s (−e 4πi/3 + e 2πi/3 )s −2/3 ds2πi 0√3i 3−2/3∫ ∞=e −s s −2/3 ds2πi= 3−1/62π Γ( 1 3 ),00e −s e 2πi/3 (3s) −2/3 dswhere we have used the st<strong>and</strong>ard integral defining the gamma function in the last line. ◭Finally in this subsection we should mention that the Airy functions <strong>and</strong> theirderivatives are closely related to Bessel functions of orders ± 1 3 <strong>and</strong> ± 2 3<strong>and</strong> that894


25.7 WKB METHODSthere exist many representations, both as linear combinations <strong>and</strong> as indefiniteintegrals, of one in terms of the other. §25.7 WKB methodsThroughout this book we have had many occasions on which it has been necessaryto solve the equationd 2 ydx 2 + k2 0f(x)y = 0 (25.42)when the notionally general function f(x) has been, in fact, a constant, usuallythe unit function f(x) = 1. Then the solutions have been elementary <strong>and</strong> of the<strong>for</strong>m A sin k 0 x or A cos k 0 x with arbitrary but constant amplitude A.Explicit solutions of (25.42) <strong>for</strong> a non-constant f(x) are only possible in alimited number of cases, but, as we will show, some progress can be made if f(x)is a slowly varying function of x, in the sense that it does not change much in arange of x of the order of k0 −1.We will also see that it is possible to h<strong>and</strong>le situations in which f(x) iscomplex; this enables us to deal with, <strong>for</strong> example, the passage of waves throughan absorbing medium. Developing such solutions will involve us in finding theintegrals of some complex quantities, integrals that will behave differently in thevarious parts of the complex plane – hence their inclusion in this chapter.25.7.1 Phase memoryBe<strong>for</strong>e moving on to the <strong>for</strong>mal development of WKB methods we discuss theconcept of phase memory which is the underlying idea behind them.Let us first suppose that f(x) is real, positive <strong>and</strong> essentially constant overarangeofx <strong>and</strong> define n(x) as the positive square root of f(x); n(x) isthenalso real, positive <strong>and</strong> essentially constant over the same range of x. We adoptthis notation so that the connection can be made with the description of anelectromagnetic wave travelling through a medium of dielectric constant f(x)<strong>and</strong>, consequently, refractive index n(x). The quantity y(x) would be the electricor magnetic field of the wave. For this simplified case, in which we can omit the§ These relationships <strong>and</strong> many other properties of the Airy functions can be found in, <strong>for</strong> example,M. Abramowitz <strong>and</strong> I. A. Stegun (eds), H<strong>and</strong>book of <strong>Mathematical</strong> Functions (New York: Dover,1965) pp. 446–50. So called because they were used, independently, by Wentzel, Kramers <strong>and</strong> Brillouin to tacklecertain wave-mechanical problems in 1926, though they had earlier been studied in some depth byJeffreys <strong>and</strong> used as far back as the first half of the nineteenth century by Green.895


APPLICATIONS OF COMPLEX VARIABLESx-dependence of n(x), the solution would be (as usual)y(x) =A exp(−ik 0 nx), (25.43)with both A <strong>and</strong> n constant. The quantity k 0 nx would be real <strong>and</strong> would be calledthe ‘phase’ of the wave; it increases linearly with x.As a first variation on this simple picture, we may allow f(x) to be complex,though, <strong>for</strong> the moment, still constant. Then n(x) is still a constant, albeit acomplex one:n = µ + iν.The solution is <strong>for</strong>mally the same as be<strong>for</strong>e; however, whilst it still exhibitsoscillatory behaviour, the amplitude of the oscillations either grows or declines,depending upon the sign of ν:y(x) =A exp(−ik 0 nx) =A exp[ −ik 0 (µ + iν)x ]=A exp(k 0 νx) exp(−ik 0 µx).This solution with ν negative is the appropriate description <strong>for</strong> a wave travellingin a uni<strong>for</strong>m absorbing medium. The quantity k 0 (µ + iν)x is usually called thecomplex phase of the wave.We now allow f(x), <strong>and</strong> hence n(x), to be both complex <strong>and</strong> varying withposition, though, as we have noted earlier, there will be restrictions on howrapidly f(x) may vary if valid solutions are to be obtained. The obvious extensionof solution (25.43) to the present case would bey(x) =A exp[ −ik 0 n(x)x ], (25.44)but direct substitution of this into equation (25.42) givesy ′′ + k 2 0n 2 y = −k 2 0(n ′2 x 2 +2nn ′ x)y − ik 0 (n ′′ x +2n ′ )y.Clearly the RHS can only be zero, as is required by the equation, if n ′ <strong>and</strong> n ′′ areboth very small, or if some unlikely relationship exists between them.To try to improve on this situation, we consider how the phase φ of the solutionchanges as the wave passes through an infinitesimal thickness dx of the medium.The infinitesimal (complex) phase change dφ <strong>for</strong> this is clearly k 0 n(x) dx, <strong>and</strong>there<strong>for</strong>e will be∫ x∆φ = k 0 n(u) du<strong>for</strong> a finite thickness x of the medium. This suggests that an improvement on(25.44) might be∫ x)y(x) =A exp(−ik 0 n(u) du . (25.45)0This is still not an exact solution as nowy ′′ (x)+k 2 0n 2 (x)y(x) =−ik 0 n ′ (x)y(x).0896


25.7 WKB METHODSThis still requires k 0 n ′ (x) to be small (compared with, say, k0 2n2 (x)), but is someimprovement (not least in complexity!) on (25.44) <strong>and</strong> gives some measure of theconditions under which the solution might be a suitable approximation.The integral in equation (25.45) embodies what is sometimes referred to as thephase memory approach; it expresses the notion that the phase of the wave-likesolution is the cumulative effect of changes it undergoes as it passes through themedium. If the medium were uni<strong>for</strong>m the overall change would be proportionalto nx, as in (25.43); the extent to which it is not uni<strong>for</strong>m is reflected in the amountby which the integral differs from nx.The condition <strong>for</strong> solution (25.45) to be a reasonable approximation can bewritten as n ′ k0 −1 ≪ n 2 or, in words, the change in n over an x-range of k0 −1 shouldbe small compared with n 2 . For light in an optical medium, this means that therefractive index n, which is of the order of unity, must change very little over adistance of a few wavelengths.For some purposes the above approximation is adequate, but <strong>for</strong> others furtherrefinement is needed. This comes from considering solutions that are still wavelikebut have amplitudes, as well as phases, that vary with position. These are theWKB solutions developed <strong>and</strong> studied in the next three subsections.25.7.2 Constructing the WKB solutionsHaving <strong>for</strong>mulated the notion of phase memory, we now construct the WKBsolutions of the general equation (25.42), in which f(x) can now be both positiondependent<strong>and</strong> complex. As we have already seen, it is the possibility of a complexphase that permits the existence of wave-like solutions with varying amplitudes.Since n(x) is calculated as the square root of f(x), there is an ambiguity in itsoverall sign. In physical applications this is normally resolved unambiguously byconsiderations such as the inevitable increase in entropy of the system, but, so faras dealing with purely mathematical questions is concerned, the ambiguity mustbe borne in mind.The process we adopt is an iterative one based on the assumption that thesecond derivative of the complex phase with respect to x is very small <strong>and</strong> canbe approximated at each stage of the iteration. So we start with equation (25.42)<strong>and</strong> look <strong>for</strong> a solution of the <strong>for</strong>my(x) =A exp[ iφ(x)], (25.46)where A is a constant. When this is substituted into (25.42) the equation becomes[ ( ) ]2 dφ− + i d2 φdx dx 2 + k2 0n 2 (x) y(x) =0. (25.47)Setting the quantity in square brackets to zero produces a non-linear equation <strong>for</strong>897


APPLICATIONS OF COMPLEX VARIABLESwhich there is no obvious solution <strong>for</strong> a general n(x). However, on the assumptionthat d 2 φ/dx 2 is small, an iterative solution can be found.As a first approximation φ ′′ is ignored, <strong>and</strong> the solutiondφdx ≈±k 0n(x)is obtained. From this, differentiation gives an approximate value <strong>for</strong>d 2 φdx 2 ≈±k dn0dx ,which can be substituted into equation (25.47) to give, as a second approximation<strong>for</strong> dφ/dx, the expression[] 1/2dφdx ≈± k0n 2 2 dn(x) ± ik 0dx(= ± k 0 n 1 ± i dn···)2k 0 n 2 dx +≈±k 0 n + i dn2n dx .This can now be integrated to give an approximate expression <strong>for</strong> φ(x) as follows:∫ xφ(x) =± k 0 n(u) du + i ln[ n(x)], (25.48)x 02where the constant of integration has been <strong>for</strong>mally incorporated into the lowerlimit x 0 of the integral. Now, noting that exp(i 1 2 i ln n) =n−1/2 , substitution of(25.48) into equation (25.46) givesy ± (x) =A [ ∫ x]n exp ± ik 1/2 0 n(u) du(25.49)x 0as two independent WKB solutions of the original equation (25.42). This result isessentially the same as that in (25.45) except that the amplitude has been dividedby √ n(x), i.e. by [ f(x)] 1/4 .Sincef(x) may be complex, this may introduce anadditional x-dependent phase into the solution as well as the more obvious changein amplitude.◮Find two independent WKB solutions of Stokes’ equation in the <strong>for</strong>md 2 y+ λxy =0, with λ real <strong>and</strong> > 0.dx2 The <strong>for</strong>m of the equation is the same as that in (25.42) with f(x) =x, <strong>and</strong>there<strong>for</strong>en(x) =x 1/2 . The WKB solutions can be read off immediately using (25.49), so long aswe remember that although f(x) is real, it has four fourth roots <strong>and</strong> that there<strong>for</strong>e theconstant appearing in a solution can be complex. Two independent WKB solutions arey ± (x) = A ±|x| 1/4 exp [± i √ λ∫ x√ udu]= A ±|x| 1/4 exp [± i 2√ λ3 x3/2 ].898(25.50)


25.7 WKB METHODSThe precise combination of these two solutions that is required <strong>for</strong> any particular problemhas to be determined from the problem. ◭When Stokes’ equation is applied more generally to functions of a complexvariable, i.e. the real variable x is replaced by the complex variable z, ithassolutions whose type of behaviour depends upon where z lies in the complexplane. For the particular case λ = −1, when Stokes’ equation takes the <strong>for</strong>md 2 ydz 2 = zy<strong>and</strong> the two WKB solutions (with the inverse fourth root written explicitly) arey 1,2 (z) = A [1,2z exp ∓ 2 ]1/4 3 z3/2 , (25.51)one of the solutions, Ai(z) (see section 25.6), has the property that it is realwhenever z is real, whether positive or negative. For negative real z it hassinusoidal behaviour, but it becomes an evanescent wave <strong>for</strong> real positive z.Since the function z 3/2 has a branch point at z = 0 <strong>and</strong> there<strong>for</strong>e has an abrupt(complex) change in its argument there, it is clear that neither of the two functionsin (25.51), nor any fixed combination of them, can be equal to Ai(z) <strong>for</strong> all valuesof z. More explicitly, <strong>for</strong> z real <strong>and</strong> positive, Ai(z) is proportional to y 1 (z), whichis real <strong>and</strong> has the <strong>for</strong>m of a decaying exponential function, whilst <strong>for</strong> z real <strong>and</strong>negative, when z 3/2 is purely imaginary <strong>and</strong> y 1 (z) <strong>and</strong>y 2 (z) are both oscillatory,it is clear that Ai(z) must contain both y 1 <strong>and</strong> y 2 with equal amplitudes.The actual combinations of y 1 (z) <strong>and</strong>y 2 (z) needed to coincide with these twoasymptotic <strong>for</strong>ms of Ai(z) are as follows.[1For z real <strong>and</strong> > 0, c 1 y 1 (z) =2 √ πz exp − 2 ]1/4 3 z3/2 . (25.52)For z real <strong>and</strong> < 0, c 2 [ y 1 (z)e iπ/4 − y 2 (z)e −iπ/4 ]=1√ π(−z)1/4 sin [ 23 (−z)3/2 + π 4]. (25.53)There<strong>for</strong>e it must be the case that the constants used to <strong>for</strong>m Ai(z) fromthe solutions (25.51) change as z moves from one part of the complex plane toanother. In fact, the changes occur <strong>for</strong> particular values of the argument of z;these boundaries are there<strong>for</strong>e radial lines in the complex plane <strong>and</strong> are knownas Stokes lines. For Stokes’ equation they occur when arg z is equal to 0, 2π/3 or4π/3.The general occurrence of a change in the arbitrary constants used to makeup a solution, as its argument crosses certain boundaries in the complex plane, isknown as the Stokes phenomenon <strong>and</strong> is discussed further in subsection 25.7.4.899


APPLICATIONS OF COMPLEX VARIABLES◮Apply the WKB method to the problem of finding the quantum energy levels E of aparticle of mass m bound in a symmetrical one-dimensional potential well V (x) that hasonly a single minimum. The relevant Schrödinger equation is− 2 d 2 ψ+ V (x)ψ = Eψ.2m dx2 Relate the problem close to each of the classical ‘turning points’, x = ± a at which E −V (x) =0, to Stokes’ equation <strong>and</strong> assume that it is appropriate to use the solution Ai(x)given in equations (25.52) <strong>and</strong> (25.53) at x = a. Show that if the general WKB solution inthe ‘classically allowed’ region −a


25.7 WKB METHODS<strong>for</strong> some constant A. This <strong>for</strong>m is only valid <strong>for</strong> negative z close to z = 0 <strong>and</strong> is notappropriate within the well as a whole, where the approximation (25.55) leading to Stokes’equation is not valid. However, it does allow us to determine the correct combination ofthe WKB solutions found earlier <strong>for</strong> the proper continuation inside the well of the solutionfound <strong>for</strong> z>0. This isψ 1 (x) = √ A (∫ asin k(u) du + π ).k(x) x 4A similar argument gives the continuation inside the well of the evanescent solutionrequired in the region x 0 in the range −a


APPLICATIONS OF COMPLEX VARIABLESThe result just proved gives∫ an√ 2mV0−a n(a 2sn − x 2s ) 1/2 dx =(n + 1 )π. 2Writing x = va n shows that the integral is proportional to a s+1n I s ,whereI s is the integralbetween −1 <strong>and</strong> +1 of (1 − v 2s ) 1/2 <strong>and</strong> does not depend upon n. Thus E n ∝ a 2sn <strong>and</strong>a s+1n ∝ (n + 1 ), implying that E 2 n ∝ (n + 1 2 )2s/s+1 .Although not asked <strong>for</strong>, we note that the above result indicates that, <strong>for</strong> a simpleharmonic oscillator, <strong>for</strong> which s = 1, the energy levels [ E n ∼ (n + 1 ) ] are equally spaced,2whilst <strong>for</strong> very large s, corresponding to a square well, the energy levels vary as n 2 .Bothof these results agree with what is found from detailed analyses of the individual cases. ◭25.7.3 Accuracy of the WKB solutionsWe may also ask when we can expect the WKB solutions to the Stokes’ equationto be reasonable approximations. Although our final <strong>for</strong>m <strong>for</strong> the WKB solutionsis not exactly that used when the condition |n ′ k0 −1|≪|n2| was derived, it shouldgive the same order of magnitude restriction as a more careful analysis. For thederivation of (25.51), k0 2 = −1, n(z) =[f(z)]1/2 = z 1/2 , <strong>and</strong> the criterion becomes12 |z−1/2 |≪|z|, or, in round terms, |z| 3 ≫ 1.For the more general equation, typified by (25.42), the condition <strong>for</strong> the validityof the WKB solutions can usually be satisfied by making some quantity, often |z|,sufficiently large. Alternatively, a parameter such as k 0 can be made large enoughthat the validity criterion is satisfied to any pre-specified level. However, from apractical point of view, natural physical parameters cannot be varied at will, <strong>and</strong>requiring z to be large may well reduce the value of the method to virtually zero.It is normally more useful to try to obtain an improvement on a WKB solutionby multiplying it by a series whose terms contain increasing inverse powers ofthe variable, so that the result can be applied successfully <strong>for</strong> moderate, <strong>and</strong> notjust excessively large, values of the variable.We do not have the space to discuss the properties <strong>and</strong> pitfalls of suchasymptotic expansions in any detail, but exercise 25.18 will provide the readerwith a model of the general procedure. A few particular points that should benoted are given as follows.(i) If the multiplier is analytic as z →∞, then it will be represented by aseries that is convergent <strong>for</strong> |z| greater than some radius of convergenceR.(ii) If the multiplier is not analytic as z →∞, as is usually the case, then themultiplier series eventually diverges <strong>and</strong> there is a z-dependent optimalnumber of terms that the series should contain in order to give the bestaccuracy.(iii) For a fixed value of arg z, the asymptotic expansion of the multiplier isunique. However, the same asymptotic expansion can represent more than902


25.7 WKB METHODSone function <strong>and</strong> the same function may need different expansions <strong>for</strong>different values of arg z.Finally in this subsection we note that, although the <strong>for</strong>m of equation (25.42) mayappear rather restrictive, in that it contains no term in y ′ , the results obtained sofar can be applied to an equation such asd 2 ydz 2 + P (z)dy + Q(z)y =0. (25.56)dzTo make this possible, a change of either the dependent or the independentvariable is made. For the <strong>for</strong>mer we write( ∫ 1 zY (z) =y(z)exp P (u) du)⇒d2 Y(Q2dz 2 + − 1 4 P 2 − 1 )dPY =0,2 dzwhilst <strong>for</strong> the latter we introduce a new independent variable ζ defined by( ∫dζz)( ) 2dz =exp − P (u) du ⇒d2 y dzdζ 2 + Q y =0.dζIn either case, equation (25.56) is reduced to the <strong>for</strong>m of (25.42), though itwill be clear that the two sets of WKB solutions (which are, of course, onlyapproximations) will not be the same.25.7.4 The Stokes phenomenonAs we saw in subsection 25.7.2, the combination of WKB solutions of a differentialequation required to reproduce the asymptotic <strong>for</strong>m of the accurate solution y(z)of the same equation, varies according to the region of the z-plane in which z lies.We now consider this behaviour, known as the Stokes phenomenon, in a littlemore detail.Let y 1 (z) <strong>and</strong>y 2 (z) be the two WKB solutions of a second-order differentialequation. Then any solution Y (z) of the same equation can be written asymptoticallyasY (z) ∼ A 1 y 1 (z)+A 2 y 2 (z), (25.57)where, although we will be considering (abrupt) changes in them, we will continueto refer to A 1 <strong>and</strong> A 2 as constants, as they are within any one region. In order toproduce the required change in the linear combination, as we pass over a Stokesline from one region of the z-plane to another, one of the constants must change(relative to the other) as the border between the regions is crossed.At first sight, this may seem impossible without causing a discernible discontinuityin the representation of Y (z). However, we must recall that the WKBsolutions are approximations, <strong>and</strong> that, as they contain a phase integral, <strong>for</strong>certain values of arg z the phase φ(z) will be purely imaginary <strong>and</strong> the factors903


APPLICATIONS OF COMPLEX VARIABLESexp[± iφ(z) ] will be purely real. What is more, one such factor, known as thedominant term, will be exponentially large, whilst the other (the subdominant term)will be exponentially small. A Stokes line is precisely where this happens.We can now see how the change takes place without an observable discontinuityoccurring. Suppose that y 1 (z) is very large <strong>and</strong> y 2 (z) is very small on a Stokes line.Then a finite change in A 2 will have a negligible effect on Y (z); in fact, Stokesshowed, <strong>for</strong> some particular cases, that the change is less than the uncertaintyin y 1 (z) arising from the approximations made in deriving it. Since the solutionwith any particular asymptotic <strong>for</strong>m is determined in a region bounded by twoStokes lines to within an overall multiplicative constant <strong>and</strong> the original equationis linear, the change in A 2 when one of the Stokes lines is crossed must beproportional to A 1 ,i.e.A 2 changes to A 2 + SA 1 ,whereS is a constant (the Stokesconstant) characteristic of the particular line but independent of A 1 <strong>and</strong> A 2 .Itshould be emphasised that, at a Stokes line, if the dominant term is not presentin a solution, then the multiplicative constant in the subdominant term cannotchange as the line is crossed.As an example, consider the Bessel function J 0 (z) ofzeroorder.Itissinglevalued,differentiable everywhere, <strong>and</strong> can be written as a series in powers of z 2 .Itis there<strong>for</strong>e an integral even function of z. However, its asymptotic approximations<strong>for</strong> two regions of the z-plane, Re z>0<strong>and</strong>z real <strong>and</strong> negative, are given byJ 0 (z) ∼ √ 1 1 √z(e iz e −iπ/4 + e −iz e iπ/4) , | arg(z)| < 12π2 π, | arg(z−1/2 )| < 1 4 π,J 0 (z) ∼ 1 √2π1 √z(e iz e 3iπ/4 + e −iz e iπ/4) , arg(z) =π, arg(z −1/2 )=− 1 2 π.We note in passing that neither of these expressions is naturally single-valued,<strong>and</strong> a prescription <strong>for</strong> taking the square root has to be given. Equally, neither isan even function of z. For our present purpose the important point to note isthat, <strong>for</strong> both expressions, on the line arg z = π/2 both z-dependent exponentsbecome real. For large |z| the second term in each expression is large; this is thedominant term, <strong>and</strong> its multiplying constant e iπ/4 is the same in both expressions.Contrarywise, the first term in each expression is small, <strong>and</strong> its multiplyingconstant does change, from e −iπ/4 to e 3iπ/4 ,asargz passes through π/2 whilstincreasing from 0 to π. It is straight<strong>for</strong>ward to calculate the Stokes constant <strong>for</strong>this Stokes line as follows:S = A 2(new) − A 2 (old)A 1= e3iπ/4 − e −iπ/4e iπ/4= e iπ/2 − e −iπ/2 =2i.If we had moved (in the negative sense) from arg z = 0 to arg z = −π, the relevantStokes line would have been arg z = −π/2. There the first term in each expressionis dominant, <strong>and</strong> it would have been the constant e iπ/4 in the second term thatwould have changed. The final argument of z −1/2 would have been +π/2.904


25.8 APPROXIMATIONS TO INTEGRALSFinally, we should mention that the lines in the z-plane on which the exponentsin the WKB solutions are purely imaginary, <strong>and</strong> the two solutions haveequal amplitudes, are usually called the anti-Stokes lines. For the general Bessel’sequation they are the real positive <strong>and</strong> real negative axes.25.8 Approximations to integralsIn this section we will investigate a method of finding approximations to thevalues or <strong>for</strong>ms of certain types of infinite integrals. The class of integrals tobe considered is that containing integr<strong>and</strong>s that are, or can be, represented byexponential functions of the general <strong>for</strong>m g(z) exp[ f(z) ]. The exponents f(z) maybe complex, <strong>and</strong> so integrals of sinusoids can be h<strong>and</strong>led as well as those withmore obvious exponential properties. We will be using the analyticity propertiesof the functions of a complex variable to move the integration path to a partof the complex plane where a general integr<strong>and</strong> can be approximated well by ast<strong>and</strong>ard <strong>for</strong>m; the st<strong>and</strong>ard <strong>for</strong>m is then integrated explicitly.The particular st<strong>and</strong>ard <strong>for</strong>m to be employed is that of a Gausssian function ofa real variable, <strong>for</strong> which the integral between infinite limits is well known. This<strong>for</strong>m will be generated by expressing f(z) as a Taylor series expansion about apoint z 0 , at which the linear term in the expansion vanishes, i.e. where f ′ (z) =0.Then, apart from a constant multiplier, the exponential function will behave likeexp[ 1 2 f′′ (z 0 )(z − z 0 ) 2 ] <strong>and</strong>, by choosing an appropriate direction <strong>for</strong> the contourto take as it passes through the point, this can be made into a normal Gaussianfunction of a real variable <strong>and</strong> its integral may then be found.25.8.1 Level lines <strong>and</strong> saddle pointsBe<strong>for</strong>e we can discuss the method outlined above in more detail, a number ofobservations about functions of a complex variable <strong>and</strong>, in particular, about theproperties of the exponential function need to be made. For a general analyticfunction,f(z) =φ(x, y)+iψ(x, y), (25.58)of the complex variable z = x + iy, we recall that, not only do both φ <strong>and</strong> ψsatisfy Laplace’s equation, but ∇φ <strong>and</strong> ∇ψ are orthogonal. This means that thelines on which one of φ <strong>and</strong> ψ is constant are exactly the lines on which the otheris changing most rapidly.Let us apply these observations to the functionh(z) ≡ exp[ f(z) ] = exp(φ) exp(iψ), (25.59)recalling that the functions φ <strong>and</strong> ψ are themselves real. The magnitude of h(z),given by exp(φ), is constant on the lines of constant φ, which are known as the905


APPLICATIONS OF COMPLEX VARIABLESFigure 25.12 A greyscale plot with associated contours of the value of |h(z)|,where h(z) =exp[i(z 3 +6z 2 − 15z + 8)], in the neighbourhood of one of itssaddle points; darker shading corresponds to larger magnitudes. The plot alsoshows the two level lines (thick solid lines) through the saddle <strong>and</strong> part of theline of steepest descents (dashed line) passing over it. At the saddle point, theangle between the line of steepest descents <strong>and</strong> a level line is π/4.level lines of the function. It follows that the direction in which the magnitudeof h(z) changes most rapidly at any point z is in a direction perpendicular tothe level line passing through that point. This is there<strong>for</strong>e the line through z onwhich the phase of h(z), namely ψ(z), is constant. Lines of constant phase arethere<strong>for</strong>e sometimes referred to as lines of steepest descent (or steepest ascent).We further note that |h(z)| can never be negative <strong>and</strong> that neither φ nor ψcan have a finite maximum at any point at which f(z) is analytic. This latterobservation follows from the fact that at a maximum of, say, φ(x, y), both ∂ 2 φ/∂x 2<strong>and</strong> ∂ 2 φ/∂y 2 would have to be negative; if this were so, Laplace’s equation couldnot be satisfied, leading to a contradiction. A similar argument shows that aminimum of either φ or ψ is not possible wherever f(z) is analytic. A morepositive conclusion is that, since the two unmixed second partial derivatives∂ 2 φ/∂x 2 <strong>and</strong> ∂ 2 φ/∂y 2 must have opposite signs, the only possible conclusionabout a point at which ∇φ is defined <strong>and</strong> equal to zero is that the point is asaddle point of h(z). An example of a saddle point is shown as a greyscale plotin figure 25.12 <strong>and</strong>, more pictorially, in figure 5.2.906


25.8 APPROXIMATIONS TO INTEGRALSFrom the observations contained in the two previous paragraphs, we deducethat a path that follows the lines of steepest descent (or ascent) can never <strong>for</strong>ma closed loop. On such a path, φ, <strong>and</strong> hence |h(z)|, must continue to decrease(increase) until the path meets a singularity of f(z). It also follows that if a levelline of h(z) <strong>for</strong>ms a closed loop in the complex plane, then the loop must enclosea singularity of f(z). This may (if φ →∞) or may not (if φ →−∞) produce asingularity in h(z).We now turn to the study of the behaviour of h(z) at a saddle point <strong>and</strong> howthis enables us to find an approximation to the integral of h(z) alongacontourthat can be de<strong>for</strong>med to pass through the saddle point. At a saddle point z 0 ,atwhich f ′ (z 0 ) = 0, both ∇φ <strong>and</strong> ∇ψ are zero, <strong>and</strong> consequently the magnitude <strong>and</strong>phase of h(z) are both stationary. The Taylor expansion of f(z) at such a pointtakes the <strong>for</strong>mf(z) =f(z 0 )+0+ 1 2! f′′ (z 0 )(z − z 0 ) 2 +O(z − z 0 ) 3 . (25.60)We assume that f ′′ (z 0 ) ≠ 0 <strong>and</strong> write it explicitly as f ′′ (z 0 ) ≡ Ae iα , thus definingthe real quantities A <strong>and</strong> α. Ifithappensthatf ′′ (z 0 ) = 0, then two or moresaddle points coalesce <strong>and</strong> the Taylor expansion must be continued until the firstnon-vanishing term is reached; we will not consider this case further, though thegeneral method of proceeding will be apparent from what follows. If we alsoabbreviate the (in general) complex quantity f(z 0 )tof 0 , then (25.60) takes the<strong>for</strong>mf(z) =f 0 + 1 2 Aeiα (z − z 0 ) 2 +O(z − z 0 ) 3 . (25.61)To study the implications of this approximation <strong>for</strong> h(z), we write z − z 0 asρe iθ with ρ <strong>and</strong> θ both real. Then|h(z)| = | exp(f 0 )| exp[ 1 2 Aρ2 cos(2θ + α)+O(ρ 3 )]. (25.62)This shows that there are four values of θ <strong>for</strong> which |h(z)| is independent of ρ (tosecond order). These there<strong>for</strong>e correspond to two crossing level lines given by(θ = 1 2 ±12 π − α) (<strong>and</strong> θ = 1 2 ±32 π − α) . (25.63)The two level lines cross at right angles to each other. It should be noted thatthe continuations of the two level lines away from the saddle are not straightin general. At the saddle they have to satisfy (25.63), but away from it the linesmust take whatever directions are needed to make ∇φ = 0. In figure 25.12 one ofthe level lines (|h| = 1) has a continuation (y = 0) that is straight; the other doesnot <strong>and</strong> bends away from its initial direction x =1.So far as the phase of h(z) is concerned, we havearg[ h(z)]=arg(f 0 )+ 1 2 Aρ2 sin(2θ + α)+O(ρ 3 ),which shows that there are four other directions (two lines crossing at right907


APPLICATIONS OF COMPLEX VARIABLESangles) in which the phase of h(z) is independent of ρ. They make angles of π/4with the level lines through z 0 <strong>and</strong> are given byθ = − 1 2 α, θ = 1 2 (±π − α), θ = π − 1 2 α.From our previous discussion it follows that these four directions will be thelines of steepest descent (or ascent) on moving away from the saddle point. Inparticular, the two directions <strong>for</strong> which the term cos(2θ + α) in (25.62) is negativewill be the directions in which |h(z)| decreases most rapidly from its value at thesaddle point. These two directions are antiparallel, <strong>and</strong> a steepest descents pathfollowing them is a smooth locally straight line passing the saddle point. It isknown as the line of steepest descents (l.s.d.) through the saddle point. Note that‘descents’ is plural as on this line the value of |h(z)| decreases on both sides of thesaddle. This is the line which we will make the path of the contour integral ofh(z) follow. Part of a typical l.s.d. is indicated by the dashed line in figure 25.12.25.8.2 Steepest descents methodTo help underst<strong>and</strong> how an integral along the line of steepest descents can beh<strong>and</strong>led in a mechanical way, it is instructive to consider the case where thefunction f(z) =−βz 2 <strong>and</strong> h(z) =exp(−βz 2 ). The saddle point is situated atz = z 0 = 0, with f 0 = f(z 0 )=1<strong>and</strong>f ′′ (z 0 )=−2β, implying that A =2|β| <strong>and</strong>α = ±π +argβ, withthe± sign chosen to put α in the range 0 ≤ α


25.8 APPROXIMATIONS TO INTEGRALSany of the continuations to infinity of the level lines that pass through the saddle.In practical applications the start- <strong>and</strong> end-points of the path are nearly alwaysat singularities of f(z) with Re f(z) →−∞<strong>and</strong> |h(z)| →0.We now set out the complete procedure <strong>for</strong> the simplest <strong>for</strong>m of integralevaluation that uses a method of steepest descents. Extensions, such as includinghigher terms in the Taylor expansion or having to pass through more than onesaddle point in order to have appropriate termination points <strong>for</strong> the contour, canbe incorporated, but the resulting calculations tend to be long <strong>and</strong> complicated,<strong>and</strong> we do not have space to pursue them in a book such as this one.As our general integr<strong>and</strong> we take a function of the <strong>for</strong>m g(z)h(z), where, asbe<strong>for</strong>e, h(z) =exp[f(z) ]. The function g(z) should neither vary rapidly nor havezeros or singularities close to any saddle point used to evaluate the integral.Rapidly varying factors should be incorporated in the exponent, usually in the<strong>for</strong>m of a logarithm. Provided g(z) satisfies these criteria, it is sufficient to treatit as a constant multiplier when integrating, assigning to it its value at the saddlepoint, g(z 0 ).Incorporating this <strong>and</strong> retaining only the first two non-vanishing terms inequation (25.61) gives the integr<strong>and</strong> asg(z 0 ) exp(f 0 ) exp[ 1 2 Aeiα (z − z 0 ) 2 ]. (25.64)From the way in which it was defined, it follows that on the l.s.d. the imaginarypart of f(z) isconstant(=Imf 0 ) <strong>and</strong> that the final exponent in (25.64) is eitherzero (at z 0 ) or negative. We can there<strong>for</strong>e write it as −s 2 ,wheres is real. Further,since exp[ f(z)] → 0 at the start- <strong>and</strong> end-points of the contour, we must havethat s runs from −∞ to +∞, the sense of s being chosen so that it is negativeapproaching the saddle <strong>and</strong> positive when leaving it.Making this change of variable,12 Aeiα (z − z 0 ) 2 = −s 2 , with dz = ±√2A exp[ 1 2 i(π − α)]ds, (25.65)allows us to express the contribution to the integral from the neighbourhood ofthe saddle point as√ ∫ 2 ∞±g(z 0 ) exp(f 0 )A exp[ 1 2 i(π − α)] exp(−s 2 ) ds.−∞The simple saddle-point approximation assumes that this is the only contribution,<strong>and</strong> gives as the value of the contour integral∫√2πg(z) exp[ f(z)]dz = ±CA g(z 0) exp(f 0 ) exp[ 1 2 i(π − α)], (25.66)where we have used the st<strong>and</strong>ard result that ∫ ∞−∞ exp(−s2 ) ds = √ π. The overall ±909


APPLICATIONS OF COMPLEX VARIABLESsign is determined by the direction θ in the complex plane in which the distortedcontour passes through the saddle point. If − 1 2 π


25.8 APPROXIMATIONS TO INTEGRALSto the l.s.d., i.e. on the imaginary t-axis, provides some reassurance. Whether µ ispositive or negative,h(5 + iµ) = exp(50 + 10iµ − 25 − 10iµ + µ 2 )=exp(25+µ 2 ).This is greater than h(5) <strong>for</strong> all µ <strong>and</strong> increases as |µ| increases, showing thatthe integration path really does lie at a minimum of h(t) <strong>for</strong> a traversal in thisdirection.We now give a fully worked solution to a problem that could not be easilytackled by elementary means.◮ Apply the saddle-point method to the function defined by∫ ∞F(x) = 1 cos( 1 3πs3 + xs) ds0to show that its <strong>for</strong>m <strong>for</strong> large positive real x is one that tends asymptotically to zero, henceenabling F(x) to be identified with the Airy function, Ai(x).We first express the integral as an exponential function <strong>and</strong> then make the change ofvariable s = x 1/2 t to bring it into the canonical <strong>for</strong>m ∫ g(t)exp[f(t)]dt as follows:∫ ∞F(x) = 1 cos( 1 3πs3 + xs) ds0= 1 ∫ ∞exp[ i( 1 32πs3 + xs)]ds−∞= 1 ∫ ∞x 1/2 exp[ ix 3/2 ( 1 32πt3 + t)]dt.−∞We now seek to find an approximate expression <strong>for</strong> this contour integral by de<strong>for</strong>ming itspath along the real t-axis into one passing over a saddle point of the integr<strong>and</strong>. Consideredas a function of t, the multiplying factor x 1/2 /2π is a constant, <strong>and</strong> any effects due to theproximity of its zeros <strong>and</strong> singularities to any saddle point do not arise.The saddle points are situated where0=f ′ (t) =ix 3/2 (t 2 +1) ⇒ t = ±i.For reasons discussed later, we choose to use the saddle point at t = t 0 = i . At this point,f(i) =ix 3/2 (− 1 i + i) =− 2 3 3 x3/2 <strong>and</strong> Ae iα ≡ f ′′ (i) =ix 3/2 (2i) =−2x 3/2 ,<strong>and</strong> so A =2x 3/2 <strong>and</strong> α = π.Now, exp<strong>and</strong>ing f(t) around t = i by setting t = i + ρe iθ , we havef(t) =f(i)+0+ 1 2! f′′ (i)(t − i) 2 +O[(t − i) 3 ]= − 2 3 x3/2 + 1 2 2x3/2 e iπ ρ 2 e 2iθ +O(ρ 3 ).For the l.s.d. contour that crosses the saddle point we need the second term in this lastline to decrease as ρ increases. This happens if π +2θ = ±π, i.e.ifθ =0orθ = −π (or+π); thus, the l.s.d. through the saddle is orientated parallel to the real t-axis. Given theinitial contour direction, the de<strong>for</strong>med contour should approach the saddle point from thedirection θ = −π <strong>and</strong> leave it along the line θ =0.Since−π/2 < 0 ≤ π/2, the overall signof the ‘omnibus’ approximation <strong>for</strong>mula is determined as positive.911


APPLICATIONS OF COMPLEX VARIABLESFinally, putting the various values into the <strong>for</strong>mula yields( ) 1/2 2πF(x) ∼ + g(i)exp[f(i)]exp[ 1 i(π − α)]2A( ) 1/2 (2π x 1/2=+2x 3/2 2π exp − 2 )3 x3/2 exp[ 1 i(π − π)]2(1=2 √ πx exp − 2 )1/4 3 x3/2 .This is the leading term in the asymptotic expansion of F(x),which,asshowninequation(25.39), is a particular contour integral solution of Stokes’ equation. The fact that it tendsto zero in a monotonic way as x → +∞ allows it to be identified with the Airy function,Ai(x).We may ask why the saddle point at t = −i was not used. The answer to this is asfollows. Of course, any path that starts <strong>and</strong> ends in the right sectors will suffice, butif another saddle point exists close to the one used, then the Taylor expansion actuallyemployed is likely to be less effective than if there were no other saddle points or if therewere only distant ones.An investigation of the same <strong>for</strong>m as that used at t =+i shows that the saddle at t = −iis higher by a factor of exp( 4 3 x3/2 ) <strong>and</strong> that its l.s.d. is orientated parallel to the imaginaryt-axis. Thus a path that went through it would need to go via a region of largish negativeimaginary t, over the saddle at t = −i, <strong>and</strong> then, when it reached the col at t =+i, bendsharply <strong>and</strong> follow part of the same l.s.d. as considered earlier. Thus the contribution fromthe t = −i saddle would be incomplete <strong>and</strong> roughly half of that from the t =+i saddlewould still have to be included. The more serious error would come from the first of these,as, clearly, the part of the path that lies in the plane Re t = 0 is not symmetric <strong>and</strong> is farfrom Guassian-like on the side nearer the origin. The Gaussian-path approximation usedwill there<strong>for</strong>e not be a good one, <strong>and</strong>, what is more, the resulting error will be magnified bya factor exp( 4 3 x3/2 ) compared with the best estimate. So, both on the grounds of simplicity<strong>and</strong> because the effect of the other (neglected) saddle point is likely to be less severe, wechoose to use the one at t =+i. ◭25.8.3 Stationary phase methodIn the previous subsection we showed how to use the saddle points of anexponential function of a complex variable to evaluate approximately a contourintegral of that function. This was done by following the lines of steepest descentthat passed through the saddle point; these are lines on which the phase of theexponential is constant but its amplitude is varying at the maximum possiblerate <strong>for</strong> that function. We now introduce an alternative method, one that entirelyreverses the roles of amplitude <strong>and</strong> phase. To see how such an alternative approachmight work, it is useful to study how the integral of an exponential function ofa complex variablecan be represented as the sum of infinitesimal vectors in thecomplex plane.We start by studying the familiar integralI 0 =∫ ∞−∞exp(−z 2 ) dz, (25.67)912


25.8 APPROXIMATIONS TO INTEGRALSwhich we already know has the value √ π when z is real. This choice of demonstrationmodel is not accidental, but is motivated by the fact that, as we havealready shown, in the neighbourhood of a saddle point all exponential integr<strong>and</strong>scan be approximated by a Gaussian function of this <strong>for</strong>m.The same integral can also be thought of as an integral in the complex plane,in which the integration contour happens to be along the real axis. Since theintegr<strong>and</strong> is analytic, the contour could be distorted into any other that had thesame end-points, z = −∞ <strong>and</strong> z =+∞, both on the real axis.As a particular possibility, we consider an arc of a circle of radius R centredon z =0.Itiseasilyshownthatcos2θ ≥ 1+4θ/π <strong>for</strong> −π/4


APPLICATIONS OF COMPLEX VARIABLESfrom which it follows that C(∞) =S(∞) = 1 2 . Clearly, C(−∞) =S(−∞) =− 1 2 .We are now in a position to examine these two equivalent ways of evaluating I 0in∫terms of sums of infinitesmal vectors in the complex plane. When the integral∞−∞ exp(−z2 ) dz is evaluated as a real integral, or a complex one along the realz-axis, each element dz generates a vector of length exp(−z 2 ) dz in an Arg<strong>and</strong>diagram, usually called the amplitude–phase diagram <strong>for</strong> the integral. For thisintegration, whilst all vector contributions lie along the real axis, they do differ inmagnitude, starting vanishingly small, growing to a maximum length of 1 × dz,<strong>and</strong> then reducing until they are again vanishingly small. At any stage, theirvector sum (in this case, the same as their algebraic sum) is a measure of theindefinite integral∫ xI(x) = exp(−z 2 ) dz. (25.70)−∞The total length of the vector sum when x →∞is, of course, √ π, <strong>and</strong> it shouldnot be overlooked that the sum is a vector parallel to (actually coinciding with)the real axis in the amplitude–phase diagram. Formally this indicates that theintegral is real. This ‘ordinary’ view of evaluating the integral generates the sameamplitude–phase diagram as does the method of steepest descents. This is because<strong>for</strong> this particular integr<strong>and</strong> the l.s.d. never leaves the real axis.Now consider the same integral evaluated using the <strong>for</strong>m of equation (25.69).Here, each contribution, as the integration variable goes from u to u + du, isofthe <strong>for</strong>mg(u) du =cos( 1 2 πu2 ) du + i sin( 1 2 πu2 ) du.As infinitesimal vectors in the amplitude–phase diagram, all g(u) du have the samemagnitude du, but their directions change continuously. Near u =0,whereu 2is small, the change is slow <strong>and</strong> each vector element is approximately equal to√2π exp(−iπ/4) du; these contributions are all in phase <strong>and</strong> add up to a significantvector contribution in the direction θ = −π/4. This is illustrated by the centralpart of the curve in part (b) of figure 25.13, in which the amplitude–phase diagram<strong>for</strong> the ‘ordinary’ integration, discussed above, is drawn as part (a).Part (b) of the figure also shows that the vector representing the indefiniteintegral (25.70) initially (s large <strong>and</strong> negative) spirals out, in a clockwise sense,from around the point 0 + i0 in the amplitude–phase diagram <strong>and</strong> ultimately (slarge <strong>and</strong> positive) spirals in, in an anticlockwise direction, to the point √ π + i0.The total curve is called a Cornu spiral. In physical applications, such as thediffraction of light at a straight edge, the relevant limits of integration aretypically −∞ <strong>and</strong> some finite value x. Then, as can be seen, the resulting vectorsum is complex in general, with its magnitude (the distance from 0 + i0 tothepoint on the spiral corresponding to z = x) growing steadily <strong>for</strong> x0.914


25.8 APPROXIMATIONS TO INTEGRALS(a)(b)π/4β(c)√ π∫ ∞Figure 25.13 Amplitude–phase diagrams <strong>for</strong> the integral−∞ exp(−z2 ) dzusing different contours in the complex z-plane. (a) Using the real axis, as inthe steepest descents method. (b) Using the level line z = u exp(− 1 iπ) that4passes through the saddle point, as in the stationary phase method. (c) Usinga path that makes a positive angle β (< π/4) with the z-axis.The final curve, 25.13(c), shows the amplitude-phase diagram corresponding toan integration path that is along a line making a positive angle β (0


APPLICATIONS OF COMPLEX VARIABLESstationary, the magnitude of any factor, g(z), multiplying the exponential function,exp[ f(z)]∼ exp[ Ae iα (z − z 0 ) 2 ], is at least comparable to its magnitude elsewhere,then this result can be used to obtain an approximation to the value of theintegral of h(z) =g(z) exp[ f(z) ]. This is the basis of the method of stationaryphase.Returning to the behaviour of a function exp[ f(z) ] at one of its saddle points,we can now see how the considerations of the previous paragraphs can be appliedthere. We already know, from equation (25.62) <strong>and</strong> the discussion immediatelyfollowing it, that in the equationh(z) ≈ g(z 0 ) exp(f 0 )exp{ 1 2 Aρ2 [cos(2θ + α)+i sin(2θ + α)]}(25.71)the second exponent is purely imaginary on a level line, <strong>and</strong> equal to zero atthe saddle point itself. What is more, since ∇ψ = 0 at the saddle, the phase isstationary there; on one level line it is a maximum <strong>and</strong> on the other it is aminimum. As there are two level lines through a saddle point, a path on whichthe amplitude of the integr<strong>and</strong> is constant could go straight on at the saddlepoint or it could turn through a right angle. For the moment we assume that itruns continuously through the saddle.On the level line <strong>for</strong> which the phase at the saddle point is a minimum, we canwrite the phase of h(z) as approximatelyarg g(z 0 )+Imf 0 + v 2 ,where v is real, iv 2 = 1 2 Aeiα (z − z 0 ) 2 <strong>and</strong>, as previously, Ae iα = f ′′ (z 0 ). Thene iπ/4 dv = ±√A2 eiα/2 dz, (25.72)leading to an approximation to the integral of∫∫ ∞ Ah(z) dz ≈±g(z 0 ) exp(f 0 ) exp(iv )√ 2 2 exp[ i( 1 4 π − 1 2 α)]dv−∞= ± g(z 0 ) exp(f 0 ) √ π exp(iπ/4)√A2 exp[ i( 1 4 π − 1 2 α)]√2π= ±A g(z 0) exp(f 0 ) exp[ 1 2i(π − α)]. (25.73)Result (25.68) was used to obtain the second line above. The ± ambiguity is againresolved by the direction θ of the contour; it is positive if −3π/4


25.8 APPROXIMATIONS TO INTEGRALSare self-cancelling, as discussed previously. However, the ends of the contour mustbe in regions where the integr<strong>and</strong> is vanishingly small, <strong>and</strong> so at each end ofthe level line we need to add a further section of path that makes the contourterminate correctly.Fortunately, this can be done without adding to the value of the integral. Thisis because, as noted in the second paragraph following equation (25.67), far fromthe saddle the level line will be at a finite height up a ‘precipitous cliff’ thatseparates the region where the integr<strong>and</strong> grows without limit from the one whereit tends to zero. To move down the cliff-face into the zero-level valley requiresan ever smaller step the further we move away from the saddle; as the integr<strong>and</strong>is finite, the contribution to the integral is vanishingly small. In figure 25.12, thisadditional piece of path length might, <strong>for</strong> example, correspond to the infinitesimalmove from a point on the large positive x-axis (where h(z) has value 1) to a pointjust above it (where h(z) ≈ 0).Now that <strong>for</strong>mula (25.73) has been justified, we may note that it is exactlythe same as that <strong>for</strong> the method of steepest descents, equation (25.66). A similarcalculation using the level line on which the phase is a maximum also reproducesthe steepest-descents <strong>for</strong>mula. It would appear that ‘all roads lead to Rome’.However, as we explain later, some roads are more difficult than others. Wherea problem involves using more than one saddle point, if the steepest-descentsapproach is tractable, it will usually be the more straight <strong>for</strong>ward to apply.Typical amplitude-phase diagrams <strong>for</strong> an integration along a level line thatgoes straight through the saddle are shown in parts (a) <strong>and</strong> (b) of figure 25.14.The value of the integral is given, in both magnitude <strong>and</strong> phase, by the vectorv joining the initial to the final winding points <strong>and</strong>, of course, is the same inboth cases. Part (a) corresponds to the case of the phase being a minimum at thesaddle; the vector path crosses v at an angle of −π/4. When a path on which thephase at the saddle is a maximum is used, the Cornu spiral is as in part (b) ofthe figure; then the vector path crosses v at an angle of +π/4. As can be seen,the two spirals are mirror images of each other.Clearly a straight-through level line path will start <strong>and</strong> end in different zerolevelvalleys. For one that turns through a right angle at the saddle point, theend-point could be in a different valley (<strong>for</strong> a function such as exp(−z 2 ), there isonly one other) or in the same one. In the latter case the integral will give a zerovalue, unless a singularity of h(z) happens to have been enclosed by the contour.Parts (c) <strong>and</strong> (d) of figure 25.14 illustrate the phase–amplitude diagrams <strong>for</strong> thesetwo cases. In (c) the path turns through a right angle (+π/2, as it happens) at thesaddle point, but finishes up in a different valley from that in which it started. In(d) it also turns through a right angle but returns to the same valley, albeit closeto the other precipice from that near its starting point. This makes no difference<strong>and</strong> the result is zero, the two half spirals in the diagram producing resultantsthat cancel.917


APPLICATIONS OF COMPLEX VARIABLESvv(a)(b)v(c)(d)Figure 25.14 Amplitude–phase diagrams <strong>for</strong> stationary phase integration. (a)Using a straight-through path on which the phase is a minimum. (b) Usinga straight-through path on which the phase is a maximum. (c) Using a levelline that turns through +π/2 at the saddle point but starts <strong>and</strong> finishes indifferent valleys. (d) Using a level line that turns through a right angle butfinishes in the same valley as it started. In cases (a), (b) <strong>and</strong> (c) the integralvalue is represented by v (see text). In case (d) the integral has value zero.We do not have the space to consider cases with two or more saddle points,but even more care is needed with the stationary phase approach than whenusing the steepest-descents method. At a saddle point there is only one l.s.d. butthere are two level lines. If more than one saddle point is required to reach theappropriate end-point of an integration, or an intermediate zero-level valley hasto be used, then care is needed in linking the corresponding level lines in such away that the links do not make a significant, but unknown, contribution to theintegral. Yet more complications can arise if a level line through one saddle pointcrosses a line of steepest ascent through a second saddle.We conclude this section with a worked example that has direct links to thetwo preceding sections of this chapter.918


25.8 APPROXIMATIONS TO INTEGRALS◮In the worked example in subsection 25.8.2 the function∫ ∞F(x) = 1 cos( 1 3πs3 + xs) ds(∗)0was shown to have the properties associated with the Airy function, Ai(x), whenx>0. Usethe stationary phase method to show that, <strong>for</strong> x


APPLICATIONS OF COMPLEX VARIABLESAV 0 e iωt˜LCDRI RECLFigure 25.15 The inductor–capacitor–resistor network <strong>for</strong> exercise 25.1.Bwhich can also be simplified, <strong>and</strong> gives[ (12+2 √ π(−x) exp i1/4 3 (−x)3/2 − π )].4Adding the two contributions <strong>and</strong> taking the real part of the sum, though this is notnecessary here because the sum is real anyway, we obtain(22F(x) =2 √ π(−x) cos 1/4 3 (−x)3/2 − π )4(1 2= √ sin π(−x)1/43 (−x)3/2 + π ),4in agreement with the asymptotic <strong>for</strong>m given in (25.53). ◭25.9 Exercises25.1 In the method of complex impedances <strong>for</strong> a.c. circuits, an inductance L isrepresented by a complex impedance Z L = iωL <strong>and</strong> a capacitance C by Z C =1/(iωC). Kirchhoff’s circuit laws,∑I i = 0 at a node <strong>and</strong> ∑ Z i I i = ∑ V j around any closed loop,iijare then applied as if the circuit were a d.c. one.Apply this method to the a.c. bridge connected as in figure 25.15 to showthat if the resistance R is chosen as R =(L/C) 1/2 then the amplitude of thecurrent, I R , through it is independent of the angular frequency ω of the applieda.c. voltage V 0 e iωt .Determine how the phase of I R , relative to that of the voltage source, varieswith the angular frequency ω.25.2 A long straight fence made of conducting wire mesh separates two fields <strong>and</strong>st<strong>and</strong>s one metre high. Sometimes, on fine days, there is a vertical electric fieldover flat open countryside. Well away from the fence the strength of the field isE 0 . By considering the effect of the trans<strong>for</strong>mation w =(1−z 2 ) 1/2 on the real <strong>and</strong>920


25.9 EXERCISESimaginary z-axes, find the strengths of the field (a) at a point one metre directlyabove the fence, (b) at ground level one metre to the side of the fence, <strong>and</strong> (c)at a point that is level with the top of the fence but one metre to the side of it.What is the direction of the field in case (c)?25.3 For the function( ) z + cf(z) =ln ,z − cwhere c is real, show that the real part u of f is constant on a circle of radiusc cosech u centred on the point z = c coth u. Use this result to show that theelectrical capacitance per unit length of two parallel cylinders of radii a, placedwith their axes 2d apart, is proportional to [cosh −1 (d/a)] −1 .25.4 Find a complex potential in the z-plane appropriate to a physical situation inwhich the half-plane x>0, y = 0 has zero potential <strong>and</strong> the half-plane xa, s =0<strong>and</strong>thehalf-plane r


APPLICATIONS OF COMPLEX VARIABLESThen, by the principle of the argument, the number of zeros inside C is given bythe integer (2π) −1 ∆ C [arg f(z)].It can be shown that the zeros of z 4 + z + 1 lie one in each quadrant. Use theabove method to show that the zeros in the second <strong>and</strong> third quadrants have|z| < 1.25.9 Prove that∞∑ 1n 2 + 3 n + =4π.1−∞ 4 8Carry out the summation numerically, say between −4 <strong>and</strong> 4, <strong>and</strong> note howmuch of the sum comes from values near the poles of the contour integration.25.10 This exercise illustrates a method of summing some infinite series.(a) Determine the residues at all the poles of the functionπ cot πzf(z) =a 2 + z , 2where a is a positive real constant.(b) By evaluating, in two different ways, the integral I of f(z) along the straightline joining −∞ − ia/2 <strong>and</strong>+∞−ia/2, show that∞∑ 1 π coth πa= − 1a 2 + n2 2a 2a . 2 n=1(c) Deduce the value of ∑ ∞1 n−2 .25.11 By considering the integral of( ) 2 sin αz παz sin πz , α < π 2 ,around a circle of large radius, prove that∞∑(−1) m−1 sin2 mα= 1 (mα) 2 2 .m=125.12 Use the Bromwich inversion, <strong>and</strong> contours similar to that shown in figure 25.7(a),to find the functions of which the following are the Laplace trans<strong>for</strong>ms:(a) s(s 2 + b 2 ) −1 ;(b) n!(s − a) −(n+1) ,withn a positive integer <strong>and</strong> s>a;(c) a(s 2 − a 2 ) −1 ,withs>|a|.Compare your answers with those given in a table of st<strong>and</strong>ard Laplace trans<strong>for</strong>ms.25.13 Find the function f(t) whose Laplace trans<strong>for</strong>m is¯f(s) = e−s − 1+s.s 225.14 A function f(t) has the Laplace trans<strong>for</strong>mF(s) = 1 ( ) s + i2i ln ,s − ithe complex logarithm being defined by a finite branch cut running along theimaginary axis from −i to i.(a) Convince yourself that, <strong>for</strong> t>0, f(t) can be expressed as a closed contourintegral that encloses only the branch cut.922


25.9 EXERCISES(b) Calculate F(s) on either side of the branch cut, evaluate the integral <strong>and</strong>hence determine f(t).(c) Confirm that the derivative with respect to s of the Laplace trans<strong>for</strong>mintegral of your answer is the same as that given by dF/ds.25.15 Use the contour in figure 25.7(c) to show that the function with Laplace trans<strong>for</strong>ms −1/2 is (πx) −1/2 .[ For an integr<strong>and</strong> of the <strong>for</strong>m r −1/2 exp(−rx) change variable to t = r 1/2 .]25.16 Transverse vibrations of angular frequency ω on a string stretched with constanttension T are described by u(x, t) =y(x) e −iωt ,whered 2 ydx + ω2 m(x)y(x) =0.2 THere, m(x) =m 0 f(x) is the mass per unit length of the string <strong>and</strong>, in the generalcase, is a function of x. Find the first-order W.K.B. solution <strong>for</strong> y(x).Due to imperfections in its manufacturing process, a particular string hasa small periodic variation in its linear density of the <strong>for</strong>m m(x) = m 0 [1 +ɛ sin(2πx/L)], where ɛ ≪ 1. A progressive wave (i.e. one in which no energy islost) travels in the positive x-direction along the string. Show that its amplitudefluctuates by ± 1 ɛ of its value A 4 0 at x = 0 <strong>and</strong> that, to first order in ɛ, the phaseof the wave is√ɛωL m0 πx2π Tsin2 Lahead of what it would be if the string were uni<strong>for</strong>m, with m(x) =m 0 .25.17 The equation(d 2 ydz + ν + 1 2 2 − 1 )4 z2 y =0,sometimes called the Weber–Hermite equation, has solutions known as paraboliccylinder functions. Find, to within (possibly complex) multiplicative constants,the two W.K.B. solutions of this equation that are valid <strong>for</strong> large |z|. In each case,determine the leading term <strong>and</strong> show that the multiplicative correction factor isofthe<strong>for</strong>m1+O(ν 2 /z 2 ).Identify the Stokes <strong>and</strong> anti-Stokes lines <strong>for</strong> the equation. On which of theStokes lines is the W.K.B. solution that tends to zero <strong>for</strong> z large, real <strong>and</strong> negative,the dominant solution?25.18 A W.K.B. solution of Bessel’s equation of order zero,d 2 ydz + 1 dy2 z dz + y =0, (∗)valid <strong>for</strong> large |z| <strong>and</strong> −π/2 < arg z < 3π/2, is y(z) =Az −1/2 e iz .Obtainanimprovement on this by finding a multiplier of y(z) in the <strong>for</strong>m of an asymptoticexpansion in inverse powers of z as follows.(a) Substitute <strong>for</strong> y(z) in(∗) <strong>and</strong> show that the equation is satisfied to O(z −5/2 ).(b) Now replace the constant A by A(z) <strong>and</strong> find the equation that must besatisfied by A(z). Look <strong>for</strong> a solution of the <strong>for</strong>m A(z) =z ∑ σ ∞n=0 a nz −n ,where a 0 = 1. Show that σ = 0 is the only acceptable solution to the indicialequation <strong>and</strong> obtain a recurrence relation <strong>for</strong> the a n .(c) To within a (complex) constant, the expression y(z) =A(z)z −1/2 e iz is theasymptotic expansion of the Hankel function H (1)0(z). Show that it is adivergent expansion <strong>for</strong> all values of z <strong>and</strong> estimate, in terms of z, the valueof N such that ∑ Nn=0 a nz −n−1/2 e iz gives the best estimate of H (1)0 (z).923


APPLICATIONS OF COMPLEX VARIABLES25.19 The function h(z) ofthecomplexvariablez is defined by the integralh(z) =∫ i∞−i∞exp(t 2 − 2zt) dt.(a) Make a change of integration variable, t = iu, <strong>and</strong> evaluate h(z) usingast<strong>and</strong>ard integral. Is your answer valid <strong>for</strong> all finite z?(b) Evaluate the integral using the method of steepest descents, considering inparticular the cases (i) z is real <strong>and</strong> positive, (ii) z is real <strong>and</strong> negative <strong>and</strong>(iii) z is purely imaginary <strong>and</strong> equal to iβ, whereβ is real. In each case sketchthe corresponding contour in the complex t-plane.(c) Evaluate the integral <strong>for</strong> the same three cases as specified in part (b) usingthe method of stationary phases. To determine an appropriate contour thatpasses through a saddle point t = t 0 ,writet = t 0 +(u + iv) <strong>and</strong> apply thecriterion <strong>for</strong> determining a level line. Sketch the relevant contour in eachcase, indicating what freedom there is to distort it.Comment on the accuracy of the results obtained using the approximate methodsadopted in (b) <strong>and</strong> (c).25.20 Use the method of steepest descents to show that an approximate value <strong>for</strong> theintegralF(z) =∫ ∞−∞exp[ iz( 1 5 t5 + t)]dt,where z is real <strong>and</strong> positive, is( ) 1/2 2πexp(−βz) cos(βz − 1 8zπ),where β =4/(5 √ 2).25.21 The stationary phase approximation to an integral of the <strong>for</strong>mF(ν) =∫ bag(t)e iνf(t) dt, |ν| ≫1,where f(t) is a real function of t <strong>and</strong> g(t) is a slowly varying function (whencompared with the argument of the exponential), can be written as( ) 1/2 2π ∑ Ng(t n ){ [F(ν) ∼√ exp i νf(t n )+ π ( )]}|ν| An 4 sgn νf ′′ (t n ) ,n=1where the t n are the N stationary points of f(t) that lie in a


25.10 HINTS AND ANSWERSt = −i <strong>and</strong> then approaches the origin in the fourth quadrant in a curve thatis ultimately antiparallel to the positive real axis. The other contour, C 1 ,isthemirror image of this in the real axis; it is confined to the upper half-plane, passesthrough t = i <strong>and</strong> is antiparallel to the real t-axis at both of its extremities. Thecontribution to J ν (z) from the curve C k is 1 2 H(k) ν , the function H ν(k) being knownas a Hankel function.Using the method of steepest descents, establish the leading term in an asymptoticexpansion <strong>for</strong> H ν(1) <strong>for</strong> z real, large <strong>and</strong> positive. Deduce, without detailedcalculation, the corresponding result <strong>for</strong> H ν(2) . Hence establish the asymptotic<strong>for</strong>m of J ν (z) <strong>for</strong> the same range of z.25.23 Use the method of steepest descents to find an asymptotic approximation, valid<strong>for</strong> z large, real <strong>and</strong> positive, to the function defined by∫F ν (z) = exp(−iz sin t + iνt) dt,Cwhere ν is real <strong>and</strong> non-negative <strong>and</strong> C is a contour that starts at t = −π + i∞<strong>and</strong> ends at t = −i∞.25.10 Hints <strong>and</strong> answers25.1 Apply Kirchhoff’s laws to three independent loops, say ADBA, ADEA <strong>and</strong>DBED. Eliminate other currents from the equations to obtain I R = ω 0 CV 0 [(ω0 2 −ω 2 − 2iωω 0 )/(ω0 2 + ω2 )], where ω0 2 =(LC)−1 ; |I R | = ω 0 CV 0 ; the phase of I R istan −1 [(−2ωω 0 )/(ω0 2 − ω2 )].25.3 Set c coth u 1 = −d, c coth u 2 =+d, |c cosech u| = a <strong>and</strong> note that the capacitanceis proportional to (u 2 − u 1 ) −1 .25.5 ξ = constant, ellipses x 2 (a+1) −2 +y 2 (a−1) −2 = c 2 /(4a 2 ); η = constant, hyperbolaex 2 (cos α) −2 − y 2 (sin α) −2 = c 2 . The curves are the cuts −c ≤ x ≤ c, y =0<strong>and</strong>|x| ≥c, y =0.Thecurves<strong>for</strong>η =2π are the same as those <strong>for</strong> η =0.25.7 (a) For a quarter-circular contour enclosing the first quadrant, the change in theargument of the function is 0 + 8(π/2) + 0 (since y 8 + 5 = 0 has no real roots);(b) one negative real zero; a conjugate pair in the second <strong>and</strong> third quadrants,− 3 , −1 ± i.225.9 Evaluate∫π cot πz( 1 + z)( dz1+ z)2 4around a large circle centred on the origin; residue at z = −1/2 is 0; residue atz = −1/4 is4π cot(−π/4).25.11 The behaviour of the integr<strong>and</strong> <strong>for</strong> large |z| is |z| −2 exp [ (2α − π)|z| ]. The residueat z = ±m, <strong>for</strong>eachintegerm, issin 2 (mα)(−1) m /(mα) 2 . The contour contributesnothing.Required summation = [ total sum − (m = 0 term) ]/2.25.13 Note that ¯f(s) has no pole at s =0.Fort1 in the left half-plane. For 0


APPLICATIONS OF COMPLEX VARIABLES25.17 Use the binomial theorem to exp<strong>and</strong>, in inverse powers of z, both the squareroot in the exponent <strong>and</strong> the fourth root in the multiplier, working to O(z −2 ).The leading terms are y 1 (z) =Ce −z2 /4 z ν <strong>and</strong> y 2 (z) =De z2 /4 z −(ν+1) . Stokes lines:arg z =0,π/2,π,3π/2; anti-Stokes lines: arg z =(2n +1)π/4 <strong>for</strong>n =0, 1, 2, 3. y 1is dominant on arg z = π/2 or3π/2.25.19 (a) i √ πe −z2 , valid <strong>for</strong> all z, including i √ π exp(β 2 ) in case (iii).(b) The same values as in (a). The (only) saddle point, at t 0 = z, is traversed inthe direction θ =+ 1 π in all cases, though the path in the complex t-plane varies2with each case.(c) The same values as in (a). The level lines are v = ±u. In cases (i) <strong>and</strong> (ii) thecontour turns through a right angle at the saddle point.All three methods give exact answers in this case of a quadratic exponent.25.21 Saddle points at t 1 = −z <strong>and</strong> t 2 =2z with f 1 ′′ = −18z <strong>and</strong> f′′ 2 =18z.Approximation is[( π) 1/2 cos(7νz 3 − 1 π) 4+ cos(20νz3 − 1 π)]4.9zν 1+z 2 1+4z 225.23 Saddle point at t 0 =cos −1 (ν/z) is traversed in the direction θ = − 1 4 π. F ν(z) ≈(2π/z) 1/2 exp [ i(z − 1 2 νπ − 1 4 π)]. 926


26TensorsIt may seem obvious that the quantitative description of physical processes cannotdepend on the coordinate system in which they are represented. However, we mayturn this argument around: since physical results must indeed be independent ofthe choice of coordinate system, what does this imply about the nature of thequantities involved in the description of physical processes? The study of theseimplications <strong>and</strong> of the classification of physical quantities by means of them<strong>for</strong>ms the content of the present chapter.Although the concepts presented here may be applied, with little modification,to more abstract spaces (most notably the four-dimensional space–time ofspecial or general relativity), we shall restrict our attention to our familiar threedimensionalEuclidean space. This removes the need to discuss the properties ofdifferentiable manifolds <strong>and</strong> their tangent <strong>and</strong> dual spaces. The reader who isinterested in these more technical aspects of tensor calculus in general spaces,<strong>and</strong> in particular their application to general relativity, should consult one of themany excellent textbooks on the subject. §Be<strong>for</strong>e the presentation of the main development of the subject, we begin byintroducing the summation convention, which will prove very useful in writingtensor equations in a more compact <strong>for</strong>m. We then review the effects of a changeof basis in a vector space; such spaces were discussed in chapter 8. This isfollowed by an investigation of the rotation of Cartesian coordinate systems, <strong>and</strong>finally we broaden our discussion to include more general coordinate systems <strong>and</strong>trans<strong>for</strong>mations.§ For example, R. D’Inverno, Introducing Einstein’s Relativity (Ox<strong>for</strong>d: Ox<strong>for</strong>d University Press,1992); J. Foster <strong>and</strong> J. D. Nightingale, A Short Course in General Relativity (New York: Springer,2006); B. F. Schutz, A First Course in General Relativity (Cambridge; Cambridge University Press1985).927


TENSORS26.1 Some notationBe<strong>for</strong>e proceeding further, we introduce the summation convention <strong>for</strong> subscripts,since its use looms large in the work of this chapter. The convention is thatany lower-case alphabetic subscript that appears exactly twice in any term of anexpression is understood to be summed over all the values that a subscript inthat position can take (unless the contrary is specifically stated). The subscriptedquantities may appear in the numerator <strong>and</strong>/or the denominator of a term in anexpression. This naturally implies that any such pair of repeated subscripts mustoccur only in subscript positions that have the same range of values. Sometimesthe ranges of values have to be specified but usually they are apparent from thecontext.The following simple examples illustrate what is meant (in the three-dimensionalcase):(i) a i x i st<strong>and</strong>s <strong>for</strong> a 1 x 1 + a 2 x 2 + a 3 x 3 ;(ii) a ij b jk st<strong>and</strong>s <strong>for</strong> a i1 b 1k + a i2 b 2k + a i3 b 3k ;(iii) a ij b jk c k st<strong>and</strong>s <strong>for</strong> ∑ 3 ∑ 3j=1 k=1 a ijb jk c k ;(iv)(v)∂v i∂x ist<strong>and</strong>s <strong>for</strong> ∂v 1∂x 1+ ∂v 2∂x 2+ ∂v 3∂x 3;∂ 2 φst<strong>and</strong>s <strong>for</strong> ∂2 φ∂x i ∂x i ∂x 2 + ∂2 φ1∂x 2 2+ ∂2 φ∂x 2 .3Subscripts that are summed over are called dummy subscripts <strong>and</strong> the othersfree subscripts. It is worth remarking that when introducing a dummy subscriptinto an expression, care should be taken not to use one that is already present,either as a free or as a dummy subscript. For example, a ij b jk c kl cannot, <strong>and</strong> mustnot, be replaced by a ij b jj c jl or by a il b lk c kl , but could be replaced by a im b mk c klor by a im b mn c nl . Naturally, free subscripts must not be changed at all unless theworking calls <strong>for</strong> it.Furthermore, as we have done throughout this book, we will make frequentuse of the Kronecker delta δ ij , which is defined by{1 if i = j,δ ij =0 otherwise.When the summation convention has been adopted, the main use of δ ij is toreplace one subscript by another in certain expressions. Examples might include<strong>and</strong>b j δ ij = b i ,a ij δ jk = a ij δ kj = a ik . (26.1)928


26.2 CHANGE OF BASISIn the second of these the dummy index shared by both terms on the left-h<strong>and</strong>side (namely j) has been replaced by the free index carried by the Kronecker delta(namely k), <strong>and</strong> the delta symbol has disappeared. In matrix language, (26.1) canbe written as AI = A, whereA is the matrix with elements a ij <strong>and</strong> I is the unitmatrix having the same dimensions as A.In some expressions we may use the Kronecker delta to replace indices in anumber of different ways, e.g.a ij b jk δ ki = a ij b ji or a kj b jk ,where the two expressions on the RHS are totally equivalent to one another.26.2 Change of basisIn chapter 8 some attention was given to the subject of changing the basis set (orcoordinate system) in a vector space <strong>and</strong> it was shown that, under such a change,different types of quantity behave in different ways. These results are given insection 8.15, but are summarised below <strong>for</strong> convenience, using the summationconvention. Although throughout this section we will remind the reader that weare using this convention, it will simply be assumed in the remainder of thechapter.If we introduce a set of basis vectors e 1 , e 2 , e 3 into our familiar three-dimensional(vector) space, then we can describe any vector x in terms of its componentsx 1 ,x 2 ,x 3 with respect to this basis:x = x 1 e 1 + x 2 e 2 + x 3 e 3 = x i e i ,where we have used the summation convention to write the sum in a morecompact <strong>for</strong>m. If we now introduce a new basis e ′ 1 , e′ 2 , e′ 3 related to the old one bye ′ j = S ij e i (sum over i), (26.2)where the coefficient S ij is the ith component of the vector e ′ j with respect to theunprimed basis, then we may write x with respect to the new basis asx = x ′ 1e ′ 1 + x ′ 2e ′ 2 + x ′ 3e ′ 3 = x ′ ie ′ i(sum over i).If we denote the matrix with elements S ij by S, then the components x ′ i <strong>and</strong> x iin the two bases are related byx ′ i =(S −1 ) ij x j(sum over j),where, using the summation convention, there is an implicit sum over j fromj =1toj = 3. In the special case where the trans<strong>for</strong>mation is a rotation of thecoordinate axes, the trans<strong>for</strong>mation matrix S is orthogonal <strong>and</strong> we havex ′ i =(S T ) ij x j = S ji x j (sum over j). (26.3)929


TENSORSScalars behave differently under trans<strong>for</strong>mations, however, since they remainunchanged. For example, the value of the scalar product of two vectors x · y(which is just a number) is unaffected by the trans<strong>for</strong>mation from the unprimedto the primed basis. Different again is the behaviour of linear operators. If alinear operator A is represented by some matrix A in a given coordinate systemthen in the new (primed) coordinate system it is represented by a new matrix,A ′ = S −1 AS.In this chapter we develop a general <strong>for</strong>mulation to describe <strong>and</strong> classify thesedifferent types of behaviour under a change of basis (or coordinate trans<strong>for</strong>mation).In the development, the generic name tensor is introduced, <strong>and</strong> certainscalars, vectors <strong>and</strong> linear operators are described respectively as tensors of zeroth,first <strong>and</strong> second order (the order –orrank – corresponds to the number ofsubscripts needed to specify a particular element of the tensor). Tensors of third<strong>and</strong> fourth order will also occupy some of our attention.26.3 Cartesian tensorsWe begin our discussion of tensors by considering a particular class of coordinatetrans<strong>for</strong>mation – namely rotations – <strong>and</strong> we shall confine our attention strictlyto the rotation of Cartesian coordinate systems. Our object is to study the propertiesof various types of mathematical quantities, <strong>and</strong> their associated physicalinterpretations, when they are described in terms of Cartesian coordinates <strong>and</strong>the axes of the coordinate system are rigidly rotated from a basis e 1 , e 2 , e 3 (lyingalong the Ox 1 , Ox 2 <strong>and</strong> Ox 3 axes) to a new one e ′ 1 , e′ 2 , e′ 3 (lying along the Ox′ 1 ,Ox ′ 2 <strong>and</strong> Ox′ 3 axes).Since we shall be more interested in how the components of a vector or linearoperator are changed by a rotation of the axes than in the relationship betweenthe two sets of basis vectors e i <strong>and</strong> e ′ i , let us define the trans<strong>for</strong>mation matrix Las the inverse of the matrix S in (26.2). Thus, from (26.2), the components of aposition vector x, in the old <strong>and</strong> new bases respectively, are related byx ′ i = L ij x j . (26.4)Because we are considering only rigid rotations of the coordinate axes, thetrans<strong>for</strong>mation matrix L will be orthogonal, i.e. such that L −1 = L T .There<strong>for</strong>ethe inverse trans<strong>for</strong>mation is given byx i = L ji x ′ j. (26.5)The orthogonality of L also implies relations among the elements of L thatexpress the fact that LL T = L T L = I. In subscript notation they are given byL ik L jk = δ ij <strong>and</strong> L ki L kj = δ ij . (26.6)Furthermore, in terms of the basis vectors of the primed <strong>and</strong> unprimed Cartesian930


26.3 CARTESIAN TENSORSx 2x ′ 1x ′ 2θθθO x 1Figure 26.1 Rotation of Cartesian axes by an angle θ about the x 3 -axis. Thethree angles marked θ <strong>and</strong> the parallels (broken lines) to the primed axesshow how the first two equations of (26.7) are constructed.coordinate systems, the trans<strong>for</strong>mation matrix is given byL ij = e ′ i · e j .We note that the product of two rotations is also a rotation. For example,suppose that x ′ i = L ijx j <strong>and</strong> x ′′i = M ij x ′ j ; then the composite rotation is describedbyx ′′i = M ij x ′ j = M ij L jk x k =(ML) ik x k ,corresponding to the matrix ML.◮Find the trans<strong>for</strong>mation matrix L corresponding to a rotation of the coordinate axesthrough an angle θ about the e 3 -axis (or x 3 -axis), as shown in figure 26.1.Taking x as a position vector – the most obvious choice – we see from the figure thatthe components of x with respect to the new (primed) basis are given in terms of thecomponents in the old (unprimed) basis byx ′ 1 = x 1 cos θ + x 2 sin θ,x ′ 2 = −x 1 sin θ + x 2 cos θ, (26.7)x ′ 3 = x 3 .The (orthogonal) trans<strong>for</strong>mation matrix is thusThe inverse equations arein line with (26.5). ◭⎛L = ⎝cos θ sin θ 0− sin θ cos θ 00 0 1⎞⎠ .x 1 = x ′ 1 cos θ − x ′ 2 sin θ,x 2 = x ′ 1 sin θ + x ′ 2 cos θ, (26.8)x 3 = x ′ 3,931


TENSORS26.4 First- <strong>and</strong> zero-order Cartesian tensorsUsing the above example as a guide, we may consider any set of three quantitiesv i , which are directly or indirectly functions of the coordinates x i <strong>and</strong> possiblyinvolve some constants, <strong>and</strong> ask how their values are changed by any rotation ofthe Cartesian axes. The specific question to be answered is whether the specific<strong>for</strong>ms v ′ i in the new variables can be obtained from the old ones v i using (26.4),v ′ i = L ij v j . (26.9)If so, the v i are said to <strong>for</strong>m the components of a vector or first-order Cartesiantensor v. By definition, the position coordinates are themselves the componentsof such a tensor.The first-order tensor v does not change under rotation of thecoordinate axes; nevertheless, since the basis set does change, from e 1 , e 2 , e 3 toe ′ 1 , e′ 2 , v′ 3 , the components of v must also change. The changes must be such thatv = v i e i = v ie ′ ′ i (26.10)is unchanged.Since the trans<strong>for</strong>mation (26.9) is orthogonal, the components of any suchfirst-order Cartesian tensor also obey a relation that is the inverse of (26.9),v i = L ji v ′ j. (26.11)We now consider explicit examples. In order to keep the equations to reasonableproportions, the examples will be restricted to the x 1 x 2 -plane, i.e. there areno components in the x 3 -direction. Three-dimensional cases are no different inprinciple – but much longer to write out.◮Which of the following pairs (v 1 ,v 2 ) <strong>for</strong>m the components of a first-order Cartesian tensorin two dimensions?:(i) (x 2 , −x 1 ), (ii) (x 2 ,x 1 ), (iii) (x 2 1,x 2 2).We shall consider the rotation discussed in the previous example, <strong>and</strong> to save space wedenote cos θ by c <strong>and</strong> sin θ by s.(i) Here v 1 = x 2 <strong>and</strong> v 2 = −x 1 , referred to the old axes. In terms of the new coordinatesthey will be v ′ 1 = x′ 2 <strong>and</strong> v′ 2 = −x′ 1 ,i.e.v ′ 1 = x′ 2 = −sx 1 + cx 2v ′ 2 = −x′ 1 = −cx 1 − sx 2 .(26.12)Now if we start again <strong>and</strong> evaluate v 1 ′ <strong>and</strong> v′ 2 as given by (26.9) we find thatv 1 ′ = L 11v 1 + L 12 v 2 = cx 2 + s(−x 1 )v 2 ′ = L (26.13)21v 1 + L 22 v 2 = −s(x 2 )+c(−x 1 ).The expressions <strong>for</strong> v 1 ′ <strong>and</strong> v′ 2 in (26.12) <strong>and</strong> (26.13) are the same whatever the valuesof θ (i.e. <strong>for</strong> all rotations) <strong>and</strong> thus by definition (26.9) the pair (x 2 , −x 1 ) is a first-orderCartesian tensor.932


26.4 FIRST- AND ZERO-ORDER CARTESIAN TENSORS(ii) Here v 1 = x 2 <strong>and</strong> v 2 = x 1 . Following the same procedure,v ′ 1 = x ′ 2 = −sx 1 + cx 2v ′ 2 = x ′ 1 = cx 1 + sx 2 .But, by (26.9), <strong>for</strong> a Cartesian tensor we must havev ′ 1 = cv 1 + sv 2 = cx 2 + sx 1v ′ 2 =(−s)v 1 + cv 2 = −sx 2 + cx 1 .These two sets of expressions do not agree <strong>and</strong> thus the pair (x 2 ,x 1 ) is not a first-orderCartesian tensor.(iii) v 1 = x 2 1 <strong>and</strong> v 2 = x 2 2 . As in (ii) above, considering the first component alone issufficient to show that this pair is also not a first-order tensor. Evaluating v 1 ′ directly giveswhilst (26.9) requires thatwhich is quite different. ◭v ′ 1 = x ′ 12 = c 2 x 2 1 +2csx 1 x 2 + s 2 x 2 2,v ′ 1 = cv 1 + sv 2 = cx 2 1 + sx 2 2,There are many physical examples of first-order tensors (i.e. vectors) that will befamiliar to the reader. As a straight<strong>for</strong>ward one, we may take the set of Cartesiancomponents of the momentum of a particle of mass m, (mẋ 1 ,mẋ 2 ,mẋ 3 ). This settrans<strong>for</strong>ms in all essentials as (x 1 ,x 2 ,x 3 ), since the other operations involved,multiplication by a number <strong>and</strong> differentiation with respect to time, are quiteunaffected by any orthogonal trans<strong>for</strong>mation of the axes. Similarly, acceleration<strong>and</strong> <strong>for</strong>ce are represented by the components of first-order tensors.Other more complicated vectors involving the position coordinates more thanonce, such as the angular momentum of a particle of mass m, namely J =x × p = m(x × ẋ), are also first-order tensors. That this is so is less obvious incomponent <strong>for</strong>m than <strong>for</strong> the earlier examples, but may be verified by writingout the components of J explicitly or by appealing to the quotient law to bediscussed in section 26.7 <strong>and</strong> using the Cartesian tensor ɛ ijk from section 26.8.Having considered the effects of rotations on vector-like sets of quantities wemay consider quantities that are unchanged by a rotation of axes. In our previousnomenclature these have been called scalars but we may also describe them astensors of zero order. They contain only one element (<strong>for</strong>mally, the number ofsubscripts needed to identify a particular element is zero); the most obvious nontrivialexample associated with a rotation of axes is the square of the distance ofa point from the origin, r 2 = x 2 1 + x2 2 + x2 3 . In the new coordinate system it willhave the <strong>for</strong>m r ′2 = x ′ 12 + x ′ 22 + x ′ 32 , which <strong>for</strong> any rotation has the same value asx 2 1 + x2 2 + x2 3 . 933


TENSORSIn fact any scalar product of two first-order tensors (vectors) is a zero-ordertensor (scalar), as might be expected since it can be written in a coordinate-freeway as u · v.◮By considering the components of the vectors u <strong>and</strong> v with respect to two Cartesiancoordinate systems (related by a rotation), show that the scalar product u · v is invariantunder rotation.In the original (unprimed) system the scalar product is given in terms of components byu i v i (summed over i), <strong>and</strong> in the rotated (primed) system byu ′ iv ′ i = L ij u j L ik v k = L ij L ik u j v k = δ jk u j v k = u j v j ,where we have used the orthogonality relation (26.6). Since the resulting expression in therotated system is the same as that in the original system, the scalar product is indeedinvariant under rotations. ◭The above result leads directly to the identification of many physically importantquantities as zero-order tensors. Perhaps the most immediate of these isenergy, either as potential energy or as an energy density (e.g. F · dr, eE · dr, D · E,B · H, µ · B), but others, such as the angle between two directed quantities, areimportant.As mentioned in the first paragraph of this chapter, in most analyses of physicalsituations it is a scalar quantity (such as energy) that is to be determined. Suchquantities are invariant under a rotation of axes <strong>and</strong> so it is possible to work withthe most convenient set of axes <strong>and</strong> still have confidence in the results.Complementing the way in which a zero-order tensor was obtained from twofirst-order tensors, so a first-order tensor can be obtained from a zero-order tensor(i.e. a scalar). We show this by taking a specific example, that of the electric fieldE = −∇φ; this is derived from a scalar, the electrostatic potential φ, <strong>and</strong> hascomponentsE i = − ∂φ∂x i. (26.14)Clearly, E is a first-order tensor, but we may prove this more <strong>for</strong>mally byconsidering the behaviour of its components (26.14) under a rotation of thecoordinate axes, since the components of the electric field E ′ i are then given by(E i ′ = − ∂φ ) ′= − ∂φ′∂x i ∂x ′ = − ∂x j ∂φi∂x ′ = L ij E j , (26.15)i∂x jwhere (26.5) has been used to evaluate ∂x j /∂x ′ i . Now (26.15) is in the <strong>for</strong>m(26.9), thus confirming that the components of the electric field do behave as thecomponents of a first-order tensor.934


26.5 SECOND- AND HIGHER-ORDER CARTESIAN TENSORS◮If v i are the components of a first-order tensor, show that ∇ · v = ∂v i /∂x i is a zero-ordertensor.In the rotated coordinate system ∇ · v is given by( ) ′ ∂vi= ∂v′ i= ∂x j ∂∂v k(L∂x i ∂x ′ i∂x ′ ik v k )=L ij L ik ,i∂x j ∂x jsince the elements L ij are not functions of position. Using the orthogonality relation (26.6)we then find∂v i′∂x ′ i∂v k ∂v k= L ij L ik = δ jk = ∂v j.∂x j ∂x j ∂x jHence ∂v i /∂x i is invariant under rotation of the axes <strong>and</strong> is thus a zero-order tensor; thiswas to be expected since it can be written in a coordinate-free way as ∇ · v. ◭26.5 Second- <strong>and</strong> higher-order Cartesian tensorsFollowing on from scalars with no subscripts <strong>and</strong> vectors with one subscript,we turn to sets of quantities that require two subscripts to identify a particularelement of the set. Let these quantities by denoted by T ij .Taking (26.9) as a guide we define a second-order Cartesian tensor as follows:the T ij <strong>for</strong>m the components of such a tensor if, under the same conditions as<strong>for</strong> (26.9),<strong>and</strong>T ′ij = L ik L jl T kl (26.16)T ij = L ki L lj T ′ kl. (26.17)At the same time we may define a Cartesian tensor of general order as follows.The set of expressions T ij···k <strong>for</strong>m the components of a Cartesian tensor if, <strong>for</strong> allrotations of the axes of coordinates given by (26.4) <strong>and</strong> (26.5), subject to (26.6),the expressions using the new coordinates, Tij···k ′ are given byT ij···k ′ = L ip L jq ···L kr T pq···r (26.18)<strong>and</strong>T ij···k = L pi L qj ···L rk T pq···r. ′ (26.19)It is apparent that in three dimensions, an Nth-order Cartesian tensor has 3 Ncomponents.Since a second-order tensor has two subscripts, it is natural to display itscomponents in matrix <strong>for</strong>m. The notation [T ij ] is used, as well as T, to denotethe matrix having T ij as the element in the ith row <strong>and</strong> jth column. §We may think of a second-order tensor T as a geometrical entity in a similarway to that in which we viewed linear operators (which trans<strong>for</strong>m one vector into§ We can also denote the column matrix containing the elements v i of a vector by [v i ].935


TENSORSanother, without reference to any coordinate system) <strong>and</strong> consider the matrixcontaining its components as a representation of the tensor with respect to aparticular coordinate system. Moreover, the matrix T =[T ij ], containing thecomponents of a second-order tensor, behaves in the same way under orthogonaltrans<strong>for</strong>mations T ′ = LTL T as a linear operator.However, not all linear operators are second-order tensors. More specifically,the two subscripts in a second-order tensor must refer to the same coordinatesystem. In particular, this means that any linear operator that trans<strong>for</strong>ms a vectorinto a vector in a different vector space cannot be a second-order tensor. Thus,although the elements L ij of the trans<strong>for</strong>mation matrix are written with twosubscripts, they cannot be the components of a tensor since the two subscriptseach refer to a different coordinate system.As examples of sets of quantities that are readily shown to be second-ordertensors we consider the following.(i) The outer product of two vectors. Letu i <strong>and</strong> v i , i =1, 2, 3, be the componentsof two vectors u <strong>and</strong> v, <strong>and</strong> consider the set of quantities T ij defined byT ij = u i v j . (26.20)The set T ij are called the components of the the outer product of u <strong>and</strong> v. Underrotations the components T ij becomeT ij ′ = u ′ iv j ′ = L ik u k L jl v l = L ik L jl u k v l = L ik L jl T kl , (26.21)which shows that they do trans<strong>for</strong>m as the components of a second-order tensor.Use has been made in (26.21) of the fact that u i <strong>and</strong> v i are the components offirst-order tensors.The outer product of two vectors is often denoted, without reference to anycoordinate system, asT = u ⊗ v. (26.22)(This is not to be confused with the vector product of two vectors, which is itselfa vector <strong>and</strong> is discussed in chapter 7.) The expression (26.22) gives the basis towhich the components T ij of the second-order tensor refer: since u = u i e i <strong>and</strong>v = v i e i , we may write the tensor T asT = u i e i ⊗ v j e j = u i v j e i ⊗ e j = T ij e i ⊗ e j . (26.23)Moreover, as <strong>for</strong> the case of first-order tensors (see equation (26.10)) we notethat the quantities T ij ′ are the components of the same tensor T, but referred toa different coordinate system, i.e.T = T ij e i ⊗ e j = T ije ′ ′ i ⊗ e ′ j.These concepts can be extended to higher-order tensors.936


26.5 SECOND- AND HIGHER-ORDER CARTESIAN TENSORS(ii) The gradient of a vector. Suppose v i represents the components of a vector;let us consider the quantities generated by <strong>for</strong>ming the derivatives of each v i ,i =1, 2, 3, with respect to each x j , j =1, 2, 3, i.e.T ij = ∂v i.∂x jThese nine quantities <strong>for</strong>m the components of a second-order tensor, as can beseen from the fact thatT ′ij = ∂v′ i∂x ′ j= ∂(L ikv k ) ∂x l ∂v k∂x l ∂x ′ = L ik L jl = L ik L jl T kl .j∂x lIn coordinate-free language the tensor T may be written as T = ∇v <strong>and</strong> hencegives meaning to the concept of the gradient of a vector, a quantity that was notdiscussed in the chapter on vector calculus (chapter 10).A test of whether any given set of quantities <strong>for</strong>ms the components of a secondordertensor can always be made by direct substitution of the x ′ i in terms of thex i , followed by comparison with the right-h<strong>and</strong> side of (26.16). This procedure isextremely laborious, however, <strong>and</strong> it is almost always better to try to recognisethe set as being expressible in one of the <strong>for</strong>ms just considered, or to makealternative tests based on the quotient law of section 26.7 below.◮Show that the T ij given by(x2T =[T ij ]= 2−x 1 x 2−x 1 x 2 x 2 1are the components of a second-order tensor.)(26.24)Again we consider a rotation θ about the e 3 -axis. Carrying out the direct evaluation firstwe obtain, using (26.7),T 11 ′ = x ′ 2 2 = s 2 x 2 1 − 2scx 1 x 2 + c 2 x 2 2,T 12 ′ = −x ′ 1x ′ 2 = scx 2 1 +(s 2 − c 2 )x 1 x 2 − scx 2 2,T 21 ′ = −x ′ 1x ′ 2 = scx 2 1 +(s 2 − c 2 )x 1 x 2 − scx 2 2,T 22 ′ = x ′ 1 2 = c 2 x 2 1 +2scx 1 x 2 + s 2 x 2 2.Now, evaluating the right-h<strong>and</strong> side of (26.16),T 11 ′ = ccx 2 2 + cs(−x 1 x 2 )+sc(−x 1 x 2 )+ssx 2 1,T 12 ′ = c(−s)x 2 2 + cc(−x 1 x 2 )+s(−s)(−x 1 x 2 )+scx 2 1,T 21 ′ =(−s)cx 2 2 +(−s)s(−x 1 x 2 )+cc(−x 1 x 2 )+csx 2 1,T 22 ′ =(−s)(−s)x 2 2 +(−s)c(−x 1 x 2 )+c(−s)(−x 1 x 2 )+ccx 2 1.After reorganisation, the corresponding expressions are seen to be the same, showing, asrequired, that the T ij are the components of a second-order tensor.The same result could be inferred much more easily, however, by noting that the T ijare in fact the components of the outer product of the vector (x 2 , −x 1 ) with itself. That(x 2 , −x 1 ) is indeed a vector was established by (26.12) <strong>and</strong> (26.13). ◭937


TENSORSPhysical examples involving second-order tensors will be discussed in the latersections of this chapter, but we might note here that, <strong>for</strong> example, magneticsusceptibility <strong>and</strong> electrical conductivity are described by second-order tensors.26.6 The algebra of tensorsBecause of the similarity of first- <strong>and</strong> second-order tensors to column vectors <strong>and</strong>matrices, it would be expected that similar types of algebraic operation can becarried out with them <strong>and</strong> so provide ways of constructing new tensors from oldones. In the remainder of this chapter, instead of referring to the T ij (say) as thecomponents of a second-order tensor T, we may sometimes simply refer to T ijas the tensor. It should always be remembered, however, that the T ij are in factjust the components of T in a given coordinate system <strong>and</strong> that T ij ′ refers to thecomponents of the same tensor T in a different coordinate system.The addition <strong>and</strong> subtraction of tensors follows an obvious definition; namelythat if V ij···k <strong>and</strong> W ij···k are (the components of) tensors of the same order, thentheir sum <strong>and</strong> difference, S ij···k <strong>and</strong> D ij···k respectively, are given byS ij···k = V ij···k + W ij···k ,D ij···k = V ij···k − W ij···k ,<strong>for</strong> each set of values i,j,...,k.ThatS ij···k <strong>and</strong> D ij···k are the components oftensors follows immediately from the linearity of a rotation of coordinates.It is equally straight<strong>for</strong>ward to show that if the T ij···k are the components ofa tensor, then so is the set of quantities <strong>for</strong>med by interchanging the order of (apair of) indices, e.g. T ji···k .If T ji···k is found to be identical with T ij···k then T ij···k is said to be symmetricwith respect to its first two subscripts (or simply ‘symmetric’, <strong>for</strong> second-ordertensors). If, however, T ji···k = −T ij···k <strong>for</strong> every element then it is an antisymmetrictensor. An arbitrary tensor is neither symmetric nor antisymmetric but can alwaysbe written as the sum of a symmetric tensor S ij···k <strong>and</strong> an antisymmetric tensorA ij···k :T ij···k = 1 2 (T ij···k + T ji···k )+ 1 2 (T ij···k − T ji···k )= S ij···k + A ij···k .Of course these properties are valid <strong>for</strong> any pair of subscripts.In (26.20) in the previous section we had an example of a kind of ‘multiplication’of two tensors, thereby producing a tensor of higher order – in that case twofirst-order tensors were multiplied to give a second-order tensor. Inspection of(26.21) shows that there is nothing particular about the orders of the tensorsinvolved <strong>and</strong> it follows as a general result that the outer product of an Nth-ordertensor with an Mth-order tensor will produce an (M + N)th-order tensor.938


26.7 THE QUOTIENT LAWAn operation that produces the opposite effect – namely, generates a tensorof smaller rather than larger order – is known as contraction <strong>and</strong> consists ofmaking two of the subscripts equal <strong>and</strong> summing over all values of the equalisedsubscripts.◮Show that the process of contraction of an Nth-order tensor produces another tensor, o<strong>for</strong>der N − 2.Let T ij···l···m···k be the components of an Nth-order tensor, thenij···l···m···k = L ip L jq ···L lr ···L ms ···L kn T pq···r···s···n .} {{ }N factorsT ′Thus if, <strong>for</strong> example, we make the two subscripts l <strong>and</strong> m equal <strong>and</strong> sum over all valuesof these subscripts, we obtainT ij···l···l···k ′ = L ip L jq ···L lr ···L ls ···L kn T pq···r···s···n= L ip L jq ···δ rs ···L kn T pq···r···s···n= L ip L jq ···L kn} {{ }(N − 2) factorsT pq···r···r···n ,showing that T ij···l···l···k are the components of a (different) Cartesian tensor of orderN − 2. ◭For a second-rank tensor, the process of contraction is the same as taking thetrace of the corresponding matrix. The trace T ii itself is thus a zero-order tensor(or scalar) <strong>and</strong> hence invariant under rotations, as was noted in chapter 8.The process of taking the scalar product of two vectors can be recast into tensorlanguage as <strong>for</strong>ming the outer product T ij = u i v j of two first-order tensors u <strong>and</strong>v <strong>and</strong> then contracting the second-order tensor T so <strong>for</strong>med, to give T ii = u i v i ,ascalar (invariant under a rotation of axes).As yet another example of a familiar operation that is a particular case of acontraction, we may note that the multiplication of a column vector [u i ]byamatrix [B ij ] to produce another column vector [v i ],B ij u j = v i ,can be looked upon as the contraction T ijj of the third-order tensor T ijk <strong>for</strong>medfrom the outer product of B ij <strong>and</strong> u k .26.7 The quotient lawThe previous paragraph appears to give a heavy-h<strong>and</strong>ed way of describing afamiliar operation, but it leads us to ask whether it has a converse. To put thequestion in more general terms: if we know that B <strong>and</strong> C are tensors <strong>and</strong> alsothatA pq···k···m B ij···k···n = C pq···mij···n , (26.25)939


TENSORSdoes this imply that the A pq···k···m also <strong>for</strong>m the components of a tensor A? HereA, B <strong>and</strong> C are respectively of Mth, Nth <strong>and</strong> (M +N −2)th order <strong>and</strong> it should benoted that the subscript k that has been contracted may be any of the subscriptsin A <strong>and</strong> B independently.The quotient law <strong>for</strong> tensors states that if (26.25) holds in all rotated coordinateframes then the A pq···k···m do indeed <strong>for</strong>m the components of a tensor A. Toproveit <strong>for</strong> general M <strong>and</strong> N is no more difficult regarding the ideas involved than toshow it <strong>for</strong> specific M <strong>and</strong> N, but this does involve the introduction of a largenumber of subscript symbols. We will there<strong>for</strong>e take the case M = N = 2, butit will be readily apparent that the principle of the proof holds <strong>for</strong> general M<strong>and</strong> N.We thus start with (say)A pk B ik = C pi , (26.26)where B ik <strong>and</strong> C pi are arbitrary second-order tensors. Under a rotation of coordinatesthe set A pk (tensor or not) trans<strong>for</strong>ms into a new set of quantities thatwe will denote by A ′ pk. We thus obtain in succession the following steps, using(26.16), (26.17) <strong>and</strong> (26.6):A ′ pk B′ ik = C′ pi (trans<strong>for</strong>ming (26.26)),= L pq L ij C qj (since C is a tensor),= L pq L ij A ql B jl (from (26.26)),= L pq L ij A ql L mj L nl B mn ′ (since B is a tensor),= L pq L nl A ql B in ′ (since L ij L mj = δ im ).Now k on the left <strong>and</strong> n on the right are dummy subscripts <strong>and</strong> thus we maywrite(A ′ pk − L pq L kl A ql )B ′ ik =0. (26.27)Since B ik , <strong>and</strong> hence Bik ′ , is an arbitrary tensor, we must haveA ′ pk = L pq L kl A ql ,showing that the A ′ pkare given by the general <strong>for</strong>mula (26.18) <strong>and</strong> hence thatthe A pk are the components of a second-order tensor. By following an analogousargument, the same result (26.27) <strong>and</strong> deduction could be obtained if (26.26) werereplaced byA pk B ki = C pi ,i.e. the contraction being now with respect to a different pair of indices.Use of the quotient law to test whether a given set of quantities is a tensor isgenerally much more convenient than making a direct substitution. A particularway in which it is applied is by contracting the given set of quantities, having940


26.8 THE TENSORS δ ij AND ɛ ijkN subscripts, with an arbitrary Nth-order tensor (i.e. one having independentlyvariable components) <strong>and</strong> determining whether the result is a scalar.◮Use the quotient law to show that the elements of T, equation (26.24), are the componentsof a second-order tensor.The outer product x i x j is a second-order tensor. Contracting this with the T ij given in(26.24) we obtainT ij x i x j = x 2 2x 2 1 − x 1 x 2 x 1 x 2 − x 1 x 2 x 2 x 1 + x 2 1x 2 2 =0,which is clearly invariant (a zeroth-order tensor). Hence by the quotient theorem T ij mustalso be a tensor. ◭26.8 The tensors δ ij <strong>and</strong> ɛ ijkIn many places throughout this book we have encountered <strong>and</strong> used the twosubscriptquantity δ ij defined by{1 if i = j,δ ij =0 otherwise.Let us now also introduce the three-subscript Levi–Civita symbol ɛ ijk , the valueof which is given by⎧⎪⎨ +1 if i, j, k is an even permutation of 1, 2, 3,ɛ ijk = −1 if i, j, k is an odd permutation of 1, 2, 3,⎪⎩ 0 otherwise.We will now show that δ ij <strong>and</strong> ɛ ijk are respectively the components of a second<strong>and</strong>a third-order Cartesian tensor. Notice that the coordinates x i do not appearexplicitly in the components of these tensors, their components consisting entirelyof 0 <strong>and</strong> 1.In passing, we also note that ɛ ijk is totally antisymmetric, i.e. it changes signunder the interchange of any pair of subscripts. In fact ɛ ijk , or any scalar multipleof it, is the only three-subscript quantity with this property.Treating δ ij first, the proof that it is a second-order tensor is straight<strong>for</strong>wardsince if, from (26.16), we consider the equationδ ′ kl = L ki L lj δ ij = L ki L li = δ kl ,we see that the trans<strong>for</strong>mation of δ ij generates the same expression (a patternof 0’s <strong>and</strong> 1’s) as does the definition of δ ij ′ in the trans<strong>for</strong>med coordinates. Thusδ ij trans<strong>for</strong>ms according to the appropriate tensor trans<strong>for</strong>mation law <strong>and</strong> isthere<strong>for</strong>e a second-order tensor.Turning now to ɛ ijk , we have to consider the quantityɛ ′ lmn = L li L mj L nk ɛ ijk . (26.28)941


TENSORSLet us begin, however, by noting that we may use the Levi–Civita symbol towrite an expression <strong>for</strong> the determinant of a 3 × 3 matrix A,|A|ɛ lmn = A li A mj A nk ɛ ijk , (26.29)which may be shown to be equivalent to the Laplace expansion (see chapter 8). §Indeed many of the properties of determinants discussed in chapter 8 can beproved very efficiently using this expression (see exercise 26.9).◮Evaluate the determinant of the matrix⎛A = ⎝ 2 1 −33 4 01 −2 1⎞⎠ .Setting l =1,m =2<strong>and</strong>n = 3 in (26.29) we find|A| = ɛ ijk A 1i A 2j A 3k= (2)(4)(1) − (2)(0)(−2) − (1)(3)(1) + (−3)(3)(−2)+ (1)(0)(1) − (−3)(4)(1) = 35,which may be verified using the Laplace expansion method. ◭We can now show that the ɛ ijk are in fact the components of a third-order tensor.Using (26.29) with the general matrix A replaced by the specific trans<strong>for</strong>mationmatrix L, we can rewrite the RHS of (26.28) in terms of |L|ɛ ′ lmn = L li L mj L nk ɛ ijk = |L|ɛ lmn .Since L is orthogonal its determinant has the value unity, <strong>and</strong> so ɛ ′ lmn = ɛ lmn.Thus we see that ɛ ′ lmn has exactly the properties of ɛ ijk but with i, j, k replaced byl,m,n, i.e. it is the same as the expression ɛ ijk written using the new coordinates.This shows that ɛ ijk is a third-order Cartesian tensor.In addition to providing a convenient notation <strong>for</strong> the determinant of a matrix,δ ij <strong>and</strong> ɛ ijk can be used to write many of the familiar expressions of vectoralgebra <strong>and</strong> calculus as contracted tensors. For example, provided we are usingright-h<strong>and</strong>ed Cartesian coordinates, the vector product a = b × c has as itsith component a i = ɛ ijk b j c k ; this should be contrasted with the outer productT = b ⊗ c, which is a second-order tensor having the components T ij = b i c j .§ This may be readily extended to an N × N matrix A, i.e.|A|ɛ i1 i 2···i N= A i1 j 1A i2 j 2 ···A iN j Nɛ j1 j 2···j N,where ɛ i1 i 2···i Nequals 1 if i 1 i 2 ···i N is an even permutation of 1, 2,...,N <strong>and</strong> equals −1 ifitisanodd permutation; otherwise it equals zero.942


26.8 THE TENSORS δ ij AND ɛ ijk◮Write the following as contracted Cartesian tensors: a·b, ∇ 2 φ, ∇×v, ∇(∇·v), ∇×(∇×v),(a × b) · c.The corresponding (contracted) tensor expressions are readily seen to be as follows:a · b = a i b i = δ ij a i b j ,∇ 2 φ =∂2 φ ∂ 2 φ= δ ij ,∂x i ∂x i ∂x i ∂x j∂v k(∇×v) i = ɛ ijk ,∂x j[∇(∇ · v)] i =∂ ( ) ∂vj ∂ 2 v j= δ jk ,∂x i ∂x j ∂x i ∂x k( )∂ ∂v m∂ 2 v m[∇×(∇×v)] i = ɛ ijk ɛ klm = ɛ ijk ɛ klm ,∂x j ∂x l ∂x j ∂x l(a × b) · c = δ ij c i ɛ jkl a k b l = ɛ ikl c i a k b l . ◭An important relationship between the ɛ- <strong>and</strong>δ- tensors is expressed by theidentityɛ ijk ɛ klm = δ il δ jm − δ im δ jl . (26.30)To establish the validity of this identity between two fourth-order tensors (theLHS is a once-contracted sixth-order tensor) we consider the various possiblecases.The RHS of (26.30) has the values+1 if i = l <strong>and</strong> j = m ≠ i, (26.31)−1 if i = m <strong>and</strong> j = l ≠ i, (26.32)0 <strong>for</strong> any other set of subscript values i, j, l, m. (26.33)In each product on the LHS k has the same value in both factors <strong>and</strong> <strong>for</strong> anon-zero contribution none of i, l, j, m can have the same value as k. Since thereare only three values, 1, 2 <strong>and</strong> 3, that any of the subscripts may take, the onlynon-zero possibilities are i = l <strong>and</strong> j = m or vice versa but not all four subscriptsequal (since then each ɛ factor is zero, as it would be if i = j or l = m). Thisreproduces (26.33) <strong>for</strong> the LHS of (26.30) <strong>and</strong> also the conditions (26.31) <strong>and</strong>(26.32). The values in (26.31) <strong>and</strong> (26.32) are also reproduced in the LHS of(26.30) since(i) if i = l <strong>and</strong> j = m, ɛ ijk = ɛ lmk = ɛ klm <strong>and</strong>, whether ɛ ijk is +1 or −1, theproduct of the two factors is +1; <strong>and</strong>(ii) if i = m <strong>and</strong> j = l, ɛ ijk = ɛ mlk = −ɛ klm <strong>and</strong> thus the product ɛ ijk ɛ klm (nosummation) has the value −1.This concludes the establishment of identity (26.30).943


TENSORSA useful application of (26.30) is in obtaining alternative expressions <strong>for</strong> vectorquantities that arise from the vector product of a vector product.◮Obtain an alternative expression <strong>for</strong> ∇×(∇×v).As shown in the previous example, ∇×(∇×v) can be expressed in tensor <strong>for</strong>m as∂ 2 v m[∇×(∇×v)] i = ɛ ijk ɛ klm∂x j ∂x l=(δ il δ jm − δ im δ jl ) ∂2 v m∂x j ∂x l= ∂ ( ) ∂vj−∂2 v i∂x i ∂x j ∂x j ∂x j=[∇(∇ · v)] i −∇ 2 v i ,where in the second line we have used the identity (26.30). This result has alreadybeen mentioned in chapter 10 <strong>and</strong> the reader is referred there <strong>for</strong> a discussion of itsapplicability. ◭By examining the various possibilities, it is straight<strong>for</strong>ward to verify that, moregenerally,∣ δ ip δ iq δ ir ∣∣∣∣∣ɛ ijk ɛ pqr =δ jp δ jq δ jr(26.34)∣ δ kp δ kq δ kr<strong>and</strong> it is easily seen that (26.30) is a special case of this result. From (26.34) wecan derive alternative <strong>for</strong>ms of (26.30), <strong>for</strong> example,ɛ ijk ɛ ilm = δ jl δ km − δ jm δ kl . (26.35)The pattern of subscripts in these identities is most easily remembered by notingthat the subscripts on the first δ on the RHS are those that immediately follow(cyclically, if necessary) the common subscript, here i, ineachɛ-term on the LHS;the remaining combinations of j,k,l,m as subscripts in the other δ-terms on theRHS can then be filled in automatically.Contracting (26.35) by setting j = l (say) we obtain, since δ kk = 3 when usingthe summation convention,ɛ ijk ɛ ijm =3δ km − δ km =2δ km ,<strong>and</strong> by contracting once more, setting k = m, we further find thatɛ ijk ɛ ijk =6. (26.36)26.9 Isotropic tensorsIt will have been noticed that, unlike most of the tensors discussed (except <strong>for</strong>scalars), δ ij <strong>and</strong> ɛ ijk have the property that all their components have valuesthat are the same whatever rotation of axes is made, i.e. the component values944


26.9 ISOTROPIC TENSORSare independent of the trans<strong>for</strong>mation L ij . Specifically, δ 11 has the value 1 inall coordinate frames, whereas <strong>for</strong> a general second-order tensor T all we knowis that if T 11 = f 11 (x 1 ,x 2 ,x 3 )thenT 11 ′ = f 11(x ′ 1 ,x′ 2 ,x′ 3 ). Tensors with the <strong>for</strong>merproperty are called isotropic (or invariant) tensors.It is important to know the most general <strong>for</strong>m that an isotropic tensor can take,since the description of the physical properties, e.g. the conductivity, magneticsusceptibility or tensile strength, of an isotropic medium (i.e. a medium havingthe same properties whichever way it is orientated) involves an isotropic tensor.In the previous section it was shown that δ ij <strong>and</strong> ɛ ijk are second- <strong>and</strong> third-orderisotropic tensors; we will now show that, to within a scalar multiple, they are theonly such isotropic tensors.Let us begin with isotropic second-order tensors. Suppose T ij is an isotropictensor; then, by definition, <strong>for</strong> any rotation of the axes we must have thatT ij = T ij ′ = L ik L jl T kl (26.37)<strong>for</strong> each of the nine components.First consider a rotation of the axes by 2π/3 about the (1, 1, 1) direction; thistakes Ox 1 , Ox 2 , Ox 3 into Ox ′ 2 , Ox′ 3 , Ox′ 1 respectively. For this rotation L 13 =1,L 21 =1,L 32 = 1 <strong>and</strong> all other L ij = 0. This requires that T 11 = T 11 ′ = T 33.Similarly T 12 = T 12 ′ = T 31. Continuing in this way, we find:(a) T 11 = T 22 = T 33 ;(b) T 12 = T 23 = T 31 ;(c) T 21 = T 32 = T 13 .Next, consider a rotation of the axes (from their original position) by π/2about the Ox 3 -axis. In this case L 12 = −1, L 21 =1,L 33 = 1 <strong>and</strong> all other L ij =0.Amongst other relationships, we must have from (26.37) that:T 13 =(−1) × 1 × T 23 ;T 23 =1× 1 × T 13 .Hence T 13 = T 23 = 0 <strong>and</strong> there<strong>for</strong>e, by parts (b) <strong>and</strong> (c) above, each element T ij =0 except <strong>for</strong> T 11 , T 22 <strong>and</strong> T 33 , which are all the same. This shows that T ij = λδ ij .◮Show that λɛ ijk is the only isotropic third-order Cartesian tensor.The general line of attack is as above <strong>and</strong> so only a minimum of explanation will be given.T ijk = T ijk ′ = L il L jm L kn T lmn (in all, there are 27 elements).Rotate about the (1, 1, 1) direction: this is equivalent to making subscript permutations1 → 2 → 3 → 1. We find(a) T 111 = T 222 = T 333 ,(b) T 112 = T 223 = T 331 (<strong>and</strong> two similar sets),(c) T 123 = T 231 = T 312 (<strong>and</strong> a set involving odd permutations of 1, 2, 3).945


TENSORSRotate by π/2 about the Ox 3 -axis: L 12 = −1, L 21 =1,L 33 = 1, the other L ij =0.(d) T 111 =(−1) × (−1) × (−1) × T 222 = −T 222 ,(e) T 112 =(−1) × (−1) × 1 × T 221 ,(f) T 221 =1× 1 × (−1) × T 112 ,(g) T 123 =(−1) × 1 × 1 × T 213 .Relations (a) <strong>and</strong> (d) show that elements with all subscripts the same are zero. Relations(e), (f) <strong>and</strong> (b) show that all elements with repeated subscripts are zero. Relations (g) <strong>and</strong>(c) show that T 123 = T 231 = T 312 = −T 213 = −T 321 = −T 132 .In total, T ijk differs from ɛ ijk by at most a scalar factor, but since ɛ ijk (<strong>and</strong> hence λɛ ijk )has already been shown to be an isotropic tensor, T ijk must be the most general third-orderisotropic Cartesian tensor. ◭Using exactly the same procedures as those employed <strong>for</strong> δ ij <strong>and</strong> ɛ ijk ,itmaybeshown that the only isotropic first-order tensor is the trivial one with all elementszero.26.10 Improper rotations <strong>and</strong> pseudotensorsSo far we have considered rigid rotations of the coordinate axes described byan orthogonal matrix L with |L| = +1, (26.4). Strictly speaking such trans<strong>for</strong>mationsare called proper rotations. We now broaden our discussion to includetrans<strong>for</strong>mations that are still described by an orthogonal matrix L but <strong>for</strong> which|L| = −1; these are called improper rotations.This kind of trans<strong>for</strong>mation can always be considered as an inversion of thecoordinate axes through the origin represented by the equationx ′ i = −x i , (26.38)combined with a proper rotation. The trans<strong>for</strong>mation may be looked uponalternatively as one that changes an initially right-h<strong>and</strong>ed coordinate system intoa left-h<strong>and</strong>ed one; any prior or subsequent proper rotation will not change thisstate of affairs. The most obvious example of a trans<strong>for</strong>mation with |L| = −1 isthe matrix corresponding to (26.38) itself; in this case L ij = −δ ij .As we have emphasised in earlier chapters, any real physical vector v may beconsidered as a geometrical object (i.e. an arrow in space), which can be referredto independently of any coordinate system <strong>and</strong> whose direction <strong>and</strong> magnitudecannot be altered merely by describing it in terms of a different coordinate system.Thus the components of v trans<strong>for</strong>m as v ′ i = L ijv j under all rotations (proper <strong>and</strong>improper).We can define another type of object, however, whose components may alsobe labelled by a single subscript but which trans<strong>for</strong>ms as v ′ i = L ijv j under properrotations <strong>and</strong> as v ′ i = −L ij v j (note the minus sign) under improper rotations. Inthis case, the v i are not strictly the components of a true first-order Cartesiantensor but instead are said to <strong>for</strong>m the components of a first-order Cartesianpseudotensor or pseudovector.946


26.10 IMPROPER ROTATIONS AND PSEUDOTENSORSx 3x ′ 1vpv ′Ox 2Ox 1x ′ 2p ′x ′ 3Figure 26.2 The behaviour of a vector v <strong>and</strong> a pseudovector p under areflection through the origin of the coordinate system x 1 ,x 2 ,x 3 giving the newsystem x ′ 1 ,x′ 2 ,x′ 3 .It is important to realise that a pseudovector (as its name suggests) is not ageometrical object in the usual sense. In particular, it should not be consideredas a real physical arrow in space, since its direction is reversed by an impropertrans<strong>for</strong>mation of the coordinate axes (such as an inversion through the origin).This is illustrated in figure 26.2, in which the pseudovector p is shown as a brokenline to indicate that it is not a real physical vector.Corresponding to vectors <strong>and</strong> pseudovectors, zeroth-order objects may bedivided into scalars <strong>and</strong> pseudoscalars – the latter being invariant under rotationbut changing sign on reflection.We may also extend the notion of scalars <strong>and</strong> pseudoscalars, vectors <strong>and</strong> pseudovectors,to objects with two or more subscripts. For two subcripts, as definedpreviously, any quantity with components that trans<strong>for</strong>m as T ij ′ = L ikL jl T kl underall rotations (proper <strong>and</strong> improper) is called a second-order Cartesian tensor.If, however, T ij ′ = L ikL jl T kl under proper rotations but T ij ′ = −L ikL jl T kl underimproper ones (which include reflections), then the T ij are the components ofa second-order Cartesian pseudotensor. In general the components of Cartesianpseudotensors of arbitary order trans<strong>for</strong>m asT ′ij···k = |L|L il L jm ···L kn T lm···n , (26.39)where |L| is the determinant of the trans<strong>for</strong>mation matrix.For example, from (26.29) we have that|L|ɛ ijk = L il L jm L kn ɛ lmn ,947


TENSORSbut since |L| = ±1 we may rewrite this asɛ ijk = |L|L il L jm L kn ɛ lmn .From this expression, we see that although ɛ ijk behaves as a tensor under properrotations, as discussed in section 26.8, it should properly be regarded as a thirdorderCartesian pseudotensor.◮If b j <strong>and</strong> c k are the components of vectors, show that the quantities a i = ɛ ijk b j c k <strong>for</strong>mthe components of a pseudovector.In a new coordinate system we havea ′ i = ɛ ′ ijkb ′ jc ′ k= |L|L il L jm L kn ɛ lmn L jp b p L kq c q= |L|L il ɛ lmn δ mp δ nq b p c q= |L|L il ɛ lmn b m c n= |L|L il a l ,from which we see immediately that the quantities a i <strong>for</strong>m the components of a pseudovector.◭The above example is worth some further comment. If we denote the vectorswith components b j <strong>and</strong> c k by b <strong>and</strong> c respectively then, as mentioned insection 26.8, the quantities a i = ɛ ijk b j c k are the components of the real vectora = b × c, provided that we are using a right-h<strong>and</strong>ed Cartesian coordinate system.However, in a coordinate system that is left-h<strong>and</strong>ed the quantitites a ′ i = ɛ′ ijk b′ j c′ kare not the components of the physical vector a = b × c, which has, instead, thecomponents −a ′ i . It is there<strong>for</strong>e important to note the h<strong>and</strong>edness of a coordinatesystem be<strong>for</strong>e attempting to write in component <strong>for</strong>m the vector relation a = b×c(which is true without reference to any coordinate system).It is worth noting that, although pseudotensors can be useful mathematicalobjects, the description of the real physical world must usually be in terms oftensors (i.e. scalars, vectors, etc.). § For example, the temperature or density of agas must be a scalar quantity (rather than a pseudoscalar), since its value doesnot change when the coordinate system used to describe it is inverted throughthe origin. Similarly, velocity, magnetic field strength or angular momentum canonly be described by a vector, <strong>and</strong> not by a pseudovector.At this point, it may be useful to make a brief comment on the distinctionbetween active <strong>and</strong> passive trans<strong>for</strong>mations of a physical system, as this differenceoften causes confusion. In this chapter, we are concerned solely with passive trans-§ In fact the quantum-mechanical description of elementary particles, such as electrons, protons <strong>and</strong>neutrons, requires the introduction of a new kind of mathematical object called a spinor, whichisnot a scalar, vector, or more general tensor. The study of spinors, however, falls beyond the scopeof this book.948


26.11 DUAL TENSORS<strong>for</strong>mations, <strong>for</strong> which the physical system of interest is left unaltered, <strong>and</strong> onlythe coordinate system used to describe it is changed. In an active trans<strong>for</strong>mation,however, the system itself is altered.As an example, let us consider a particle of mass m that is located at a positionx relative to the origin O <strong>and</strong> hence has velocity ẋ. The angular momentum ofthe particle about O is thus J = m(x × ẋ). If we merely invert the Cartesiancoordinates used to describe this system through O, neither the magnitude nordirection of any these vectors will be changed, since they may be consideredsimply as arrows in space that are independent of the coordinates used to describethem. If, however, we per<strong>for</strong>m the analogous active trans<strong>for</strong>mation onthe system, by inverting the position vector of the particle through O, thenitis clear that the direction of particle’s velocity will also be reversed, since itis simply the time derivative of the position vector, but that the direction ofits angular momentum vector remains unaltered. This suggests that vectors canbe divided into two categories, as follows: polar vectors (such as position <strong>and</strong>velocity), which reverse direction under an active inversion of the physical systemthrough the origin, <strong>and</strong> axial vectors (such as angular momentum), whichremain unchanged. It should be emphasised that at no point in this discussionhave we used the concept of a pseudovector to describe a real physicalquantity. §26.11 Dual tensorsAlthough pseudotensors are not themselves appropriate <strong>for</strong> the description ofphysical phenomena, they are sometimes needed; <strong>for</strong> example, we may use thepseudotensor ɛ ijk to associate with every antisymmetric second-order tensor A ij(in three dimensions) a pseudovector p i given byp i = 1 2 ɛ ijkA jk ; (26.40)p i is called the dual of A ij . Thus if we denote the antisymmetric tensor A by thematrix⎛0 A 12 −A 31⎞A =[A ij ]= ⎝ −A 12 0 A 23⎠A 31 −A 23 0then the components of its dual pseudovector are (p 1 ,p 2 ,p 3 )=(A 23 ,A 31 ,A 12 ).§ The scalar product of a polar vector <strong>and</strong> an axial vector is a pseudoscalar. It was the experimentaldetection of the dependence of the angular distribution of electrons of (polar vector) momentump e emitted by polarised nuclei of (axial vector) spin J N upon the pseudoscalar quantity J N · p e thatestablished the existence of the non-conservation of parity in β-decay.949


TENSORS◮Using (26.40), show that A ij = ɛ ijk p k .By contracting both sides of (26.40) with ɛ ijk , we findɛ ijk p k = 1 ɛ 2 ijkɛ klm A lm .Using the identity (26.30) then givesɛ ijk p k = 1 (δ 2 ilδ jm − δ im δ jl )A lm= 1 (A 2 ij − A ji )= 1 (A 2 ij + A ij )=A ij ,whereinthelastlineweusethefactthatA ij = −A ji . ◭By a simple extension, we may associate a dual pseudoscalar s with everytotally antisymmetric third-rank tensor A ijk , i.e. one that is antisymmetric withrespect to the interchange of every possible pair of subscripts; s is given bys = 1 3! ɛ ijkA ijk . (26.41)Since A ijk is a totally antisymmetric three-subscript quantity, we expect it toequal some multiple of ɛ ijk (since this is the only such quantity). In fact A ijk = sɛ ijk ,as can be proved by substituting this expression into (26.41) <strong>and</strong> using (26.36).26.12 Physical applications of tensorsIn this section some physical applications of tensors will be given. First-ordertensors are familiar as vectors <strong>and</strong> so we will concentrate on second-order tensors,starting with an example taken from mechanics.Consider a collection of rigidly connected point particles of which the αth,which has mass m (α) <strong>and</strong> is positioned at r (α) with respect to an origin O, istypical. Suppose that the rigid assembly is rotating about an axis through O withangular velocity ω.The angular momentum J about O of the assembly is given byJ = ∑ (r (α) × p (α)) .αBut p (α) = m (α) ṙ (α) <strong>and</strong> ṙ (α) = ω × r (α) , <strong>for</strong> any α, <strong>and</strong> so in subscript <strong>for</strong>m thecomponents of J are given byJ i = ∑ αm (α) ɛ ijk x (α)jẋ (α)k= ∑ α= ∑ α= ∑ αm (α) ɛ ijk x (α)jɛ klm ω l x m(α)m (α) (δ il δ jm − δ im δ jl )x (α)j x m (α) ω lm (α) [ (r(α) ) 2δil − x (α)ix (α)l]ω l ≡ I il ω l , (26.42)where I il is a symmetric second-order Cartesian tensor (by the quotient rule, see950


26.12 PHYSICAL APPLICATIONS OF TENSORSsection 26.7, since J <strong>and</strong> ω are vectors). The tensor is called the inertia tensor at Oof the assembly <strong>and</strong> depends only on the distribution of masses in the assembly<strong>and</strong> not upon the direction or magnitude of ω.A more realistic situation obtains if a continuous rigid body is considered. Inthis case, m (α) must be replaced everywhere by ρ(r) dx dy dz <strong>and</strong> all summationsby integrations over the volume of the body. Written out in full in Cartesians,the inertia tensor <strong>for</strong> a continuous body would have the <strong>for</strong>m⎛ ∫(y 2 + z 2 )ρdV − ∫ xyρ dV − ∫ ⎞xzρ dVI =[I ij ]= ⎝ − ∫ ∫xyρ dV (z 2 + x 2 )ρdV − ∫ yzρ dV− ∫ xzρ dV − ∫ ⎠ ,∫yzρ dV (x 2 + y 2 )ρdVwhere ρ = ρ(x, y, z) is the mass distribution <strong>and</strong> dV st<strong>and</strong>s <strong>for</strong> dx dy dz; theintegrals are to be taken over the whole body. The diagonal elements of thistensor are called the moments of inertia <strong>and</strong> the off-diagonal elements without theminus signs are known as the products of inertia.◮Show that the kinetic energy of the rotating system is given by T = 1 2 I jlω j ω l .By an argument parallel to that already made <strong>for</strong> J, the kinetic energy is given by∑T = 1 m ( (α) ṙ (α) · ṙ (α))2= 1 2= 1 2= 1 2α∑αm (α) ɛ ijk ω j x (α)kɛ ilmω l x (α)m∑m (α) (δ jl δ km − δ jm δ kl )x (α)kx(α) m ω j ω lα∑ [ (m (α) δ ) ]jl r(α) 2− x(α)x (α) ω j ω lα= 1 I 2 jlω j ω l .Alternatively, since J j = I jl ω l we may write the kinetic energy of the rotating system asT = 1 J 2 jω j . ◭The above example shows that the kinetic energy of the rotating body can beexpressed as a scalar obtained by twice contracting ω with the inertia tensor. Italso shows that the moment of inertia of the body about a line given by the unitvector ˆn is I jlˆn j ˆn l (or ˆn T Iˆn in matrix <strong>for</strong>m).Since I (≡ I jl ) is a real symmetric second-order tensor, it has associated with itthree mutually perpendicular directions that are its principal axes <strong>and</strong> have thefollowing properties (proved in chapter 8):(i) with each axis is associated a principal moment of inertia λ µ , µ =1, 2, 3;(ii) when the rotation of the body is about one of these axes, the angularvelocity <strong>and</strong> the angular momentum are parallel <strong>and</strong> given byJ = Iω = λ µ ω,i.e. ω is an eigenvector of I with eigenvalue λ µ ;951jl


TENSORS(iii) referred to these axes as coordinate axes, the inertia tensor is diagonalwith diagonal entries λ 1 ,λ 2 ,λ 3 .Two further examples of physical quantities represented by second-order tensorsare magnetic susceptibility <strong>and</strong> electrical conductivity. In the first case we have(in st<strong>and</strong>ard notation)M i = χ ij H j , (26.43)<strong>and</strong> in the second casej i = σ ij E j . (26.44)Here M is the magnetic moment per unit volume <strong>and</strong> j the current density(current per unit perpendicular area). In both cases we have on the left-h<strong>and</strong> sidea vector <strong>and</strong> on the right-h<strong>and</strong> side the contraction of a set of quantities withanother vector. Each set of quantities must there<strong>for</strong>e <strong>for</strong>m the components of asecond-order tensor.For isotropic media M ∝ H <strong>and</strong> j ∝ E, but <strong>for</strong> anisotropic materials such ascrystals the susceptibility <strong>and</strong> conductivity may be different along different crystalaxes, making χ ij <strong>and</strong> σ ij general second-order tensors, although they are usuallysymmetric.◮The electrical conductivity σ in a crystal is measured by an observer to have componentsas shown:⎛[σ ij ]= ⎝ 1 √ ⎞√ 2 02 3 1 ⎠ . (26.45)0 1 1Show that there is one direction in the crystal along which no current can flow. Does thecurrent flow equally easily in the two perpendicular directions?The current density in the crystal is given by j i = σ ij E j ,whereσ ij ,relativetotheobserver’scoordinate system, is given by (26.45). Since [σ ij ] is a symmetric matrix, it possessesthree mutually perpendicular eigenvectors (or principal axes) with respect to which theconductivity tensor is diagonal, with diagonal entries λ 1 ,λ 2 ,λ 3 , the eigenvalues of [σ ij ].As discussed in chapter 8, the eigenvalues of [σ ij ] are given by |σ − λI| = 0. Thus werequire∣ √ ∣∣∣∣∣ 1√− λ 2 02 3− λ 10 1 1− λ ∣ =0,from which we find(1 − λ)[(3 − λ)(1 − λ) − 1] − 2(1 − λ) =0.This simplifies to give λ =0, 1, 4 so that, with respect to its principal axes, the conductivitytensor has components σ ij ′ given by ⎛[σ ij] ′ = ⎝ 4 0 0⎞0 1 0 ⎠ .0 0 0Since j ′ i = σ ′ ijE ′ j, we see immediately that along one of the principal axes there is no currentflow <strong>and</strong> along the two perpendicular directions the current flows are not equal. ◭952


26.12 PHYSICAL APPLICATIONS OF TENSORSWe can extend the idea of a second-order tensor that relates two vectors to asituation where two physical second-order tensors are related by a fourth-ordertensor. The most common occurrence of such relationships is in the theory ofelasticity. This is not the place to give a detailed account of elasticity theory,but suffice it to say that the local de<strong>for</strong>mation of an elastic body at any interiorpoint P can be described by a second-order symmetric tensor e ij called the straintensor. It is given bye ij = 1 ( ∂ui+ ∂u )j,2 ∂x j ∂x iwhere u is the displacement vector describing the strain of a small volume elementwhose unstrained position relative to the origin is x. Similarly we can describethe stress in the body at P by the second-order symmetric stress tensor p ij ;thequantity p ij is the x j -component of the stress vector acting across a plane throughP whose normal lies in the x i -direction. A generalisation of Hooke’s law thenrelates the stress <strong>and</strong> strain tensors bywhere c ijkl is a fourth-order Cartesian tensor.p ij = c ijkl e kl (26.46)◮Assuming that the most general fourth-order isotropic tensor isc ijkl = λδ ij δ kl + ηδ ik δ jl + νδ il δ jk , (26.47)find the <strong>for</strong>m of (26.46) <strong>for</strong> an isotropic medium having Young’s modulus E <strong>and</strong> Poisson’sratio σ.For an isotropic medium we must have an isotropic tensor <strong>for</strong> c ijkl , <strong>and</strong> so we assume the<strong>for</strong>m (26.47). Substituting this into (26.46) yieldsp ij = λδ ij e kk + ηe ij + νe ji .But e ij is symmetric, <strong>and</strong> if we write η + ν =2µ, then this takes the <strong>for</strong>mp ij = λe kk δ ij +2µe ij ,in which λ <strong>and</strong> µ are known as Lamé constants. It will be noted that if e ij =0<strong>for</strong>i ≠ jthen the same is true of p ij , i.e. the principal axes of the stress <strong>and</strong> strain tensors coincide.Now consider a simple tension in the x 1 -direction, i.e. p 11 = S but all other p ij =0.Then denoting e kk (summed over k) byθ we have, in addition to e ij =0<strong>for</strong>i ≠ j, thethreeequationsS = λθ +2µe 11 ,0=λθ +2µe 22 ,0=λθ +2µe 33 .Adding them givesS = θ(3λ +2µ).Substituting <strong>for</strong> θ from this into the first of the three, <strong>and</strong> recalling that Young’s modulusis defined by S = Ee 11 ,givesE asE =µ(3λ +2µ). (26.48)λ + µ953


TENSORSFurther, Poisson’s ratio is defined as σ = −e 22 /e 11 (or −e 33 /e 11 ) <strong>and</strong> is thus( ) ( )( ) 1 λθ 1 λσ =e 11 2µ = Ee11e 11 2µ 3λ +2µ = λ2(λ + µ) . (26.49)Solving (26.48) <strong>and</strong> (26.49) <strong>for</strong> λ <strong>and</strong> µ gives finallyp ij =σE(1 + σ)(1 − 2σ) e kkδ ij + E(1 + σ) e ij. ◭26.13 Integral theorems <strong>for</strong> tensorsIn chapter 11, we discussed various integral theorems involving vector <strong>and</strong> scalarfields. Most notably, we considered the divergence theorem, which states that, <strong>for</strong>any vector field a, ∫∮∇ · a dV = a · ˆn dS, (26.50)VSwhere S is the surface enclosing the volume V <strong>and</strong> ˆn is the outward-pointing unitnormal to S at each point.Writing (26.50) in subscript notation, we have∫∂a kdV = a k ˆn k dS. (26.51)V ∂x k SAlthough we shall not prove it rigorously, (26.51) can be extended in an obviousmanner to relate integrals of tensor fields, rather than just vector fields, overvolumes <strong>and</strong> surfaces, with the result∫∂T ij···k···mdV = T ij···k···m ˆn k dS.V ∂x k SThis <strong>for</strong>m of the divergence theorem <strong>for</strong> general tensors can be very useful invector calculus manipulations.◮A vector field a satisfies ∇ · a =0inside some volume V <strong>and</strong> a · ˆn =0on the boundary∫ surface S. By considering the divergence theorem applied to T ij = x i a j , show thata dV = 0.VApplying the divergence theorem to T ij = x i a j we find∫∫∮∂T ij ∂(x i a j )dV = dV = x i a j ˆn j dS =0,V ∂x j V ∂x j Ssince a j ˆn j = 0. By exp<strong>and</strong>ing the volume integral we obtain∫V∫∫∂(x i a j ) ∂x idV = a j dV +∂x j V ∂x j∫= δ ij a j dVV∫= a i dV =0,V∮∮Vx i∂a j∂x jdVwhere in going from the first to the second line we used ∂x i /∂x j = δ ij <strong>and</strong> ∂a j /∂x j =0.◭954


26.14 NON-CARTESIAN COORDINATESThe other integral theorems discussed in chapter 11 can be extended in asimilar way. For example, written in tensor notation Stokes’ theorem states that,<strong>for</strong> a vector field a i ,∫∮∂a kɛ ijk ˆn i dS = a k dx k .S ∂x j CFor a general tensor field this has the straight<strong>for</strong>ward extension∫Sɛ ijk∂T lm···k···n∂x jˆn i dS =∮CT lm···k···n dx k .26.14 Non-Cartesian coordinatesSo far we have restricted our attention to the study of tensors when they aredescribed in terms of Cartesian coordinates <strong>and</strong> the axes of coordinates are rigidlyrotated, sometimes together with an inversion of axes through the origin. In theremainder of this chapter we shall extend the concepts discussed in the previoussections by considering arbitrary coordinate trans<strong>for</strong>mations from one generalcoordinate system to another. Although this generalisation brings with it severalcomplications, we shall find that many of the properties of Cartesian tensorsare still valid <strong>for</strong> more general tensors. Be<strong>for</strong>e considering general coordinatetrans<strong>for</strong>mations, however, we begin by reminding ourselves of some properties ofgeneral curvilinear coordinates, as discussed in chapter 10.The position of an arbitrary point P in space may be expressed in terms of thethree curvilinear coordinates u 1 ,u 2 ,u 3 . We saw in chapter 10 that if r(u 1 ,u 2 ,u 3 )isthe position vector of the point P then at P there exist two sets of basis vectorse i = ∂r <strong>and</strong> ɛ i = ∇u i , (26.52)∂u iwhere i =1, 2, 3. In general, the vectors in each set neither are of unit length nor<strong>for</strong>m an orthogonal basis. However, the sets e i <strong>and</strong> ɛ i are reciprocal systems ofvectors <strong>and</strong> soe i · ɛ j = δ ij . (26.53)In the context of general tensor analysis, it is more usual to denote the secondset of vectors ɛ i in (26.52) by e i , the index being placed as a superscript todistinguish it from the (different) vector e i , which is a member of the first set in(26.52). Although this positioning of the index may seem odd (not least becauseof the possibility of confusion with powers) it <strong>for</strong>ms part of a slight modificationto the summation convention that we will adopt <strong>for</strong> the remainder of this chapter.This is as follows: any lower-case alphabetic index that appears exactly twice inany term of an expression, once as a subscript <strong>and</strong> once as a superscript, istobesummed over all the values that an index in that position can take (unless the955


TENSORScontrary is specifically stated). All other aspects of the summation conventionremain unchanged.With the introduction of superscripts, the reciprocity relation (26.53) should berewritten so that both sides of (26.54) have one subscript <strong>and</strong> one superscript, i.e.ase i · e j = δ j i . (26.54)The alternative <strong>for</strong>m of the Kronecker delta is defined in a similar way topreviously, i.e. it equals unity if i = j <strong>and</strong> is zero otherwise.For similar reasons it is usual to denote the curvilinear coordinates themselvesby u 1 ,u 2 ,u 3 , with the index raised, so thate i = ∂r∂u i <strong>and</strong> e i = ∇u i . (26.55)From the first equality we see that we may consider a superscript that appears inthe denominator of a partial derivative as a subscript.Given the two bases e i <strong>and</strong> e i , we may write a general vector a equally well interms of either basis as follows:a = a 1 e 1 + a 2 e 2 + a 3 e 3 = a i e i ;a = a 1 e 1 + a 2 e 2 + a 3 e 3 = a i e i .The a i are called the contravariant components of the vector a <strong>and</strong> the a ithe covariant components, the position of the index (either as a subscript orsuperscript) serving to distinguish between them. Similarly, we may call the e i thecovariant basis vectors <strong>and</strong> the e i the contravariant ones.◮Show that the contravariant <strong>and</strong> covariant components of a vector a are given by a i = a·e i<strong>and</strong> a i = a · e i respectively.For the contravariant components, we finda · e i = a j e j · e i = a j δj i = a i ,where we have used the reciprocity relation (26.54). Similarly, <strong>for</strong> the covariant components,a · e i = a j e j · e i = a j δ j i= a i . ◭The reason that the notion of contravariant <strong>and</strong> covariant components ofa vector (<strong>and</strong> the resulting superscript notation) was not introduced earlier isthat <strong>for</strong> Cartesian coordinate systems the two sets of basis vectors e i <strong>and</strong> e i areidentical <strong>and</strong>, hence, so are the components of a vector with respect to eitherbasis. Thus, <strong>for</strong> Cartesian coordinates, we may speak simply of the componentsof the vector <strong>and</strong> there is no need to differentiate between contravariance <strong>and</strong>covariance, or to introduce superscripts to make a distinction between them.If we consider the components of higher-order tensors in non-Cartesian coordinates,there are even more possibilities. As an example, let us consider a956


26.15 THE METRIC TENSORsecond-order tensor T. Using the outer product notation in (26.23), we may writeT in three different ways:T = T ij e i ⊗ e j = T i je i ⊗ e j = T ij e i ⊗ e j ,where T ij , T i j <strong>and</strong> T ij are called the contravariant, mixed <strong>and</strong> covariant componentsof T respectively. It is important to remember that these three sets ofquantities <strong>for</strong>m the components of the same tensor T but refer to different (tensor)bases made up from the basis vectors of the coordinate system. Again, if we areusing Cartesian coordinates then all three sets of components are identical.We may generalise the above equation to higher-order tensors. Componentscarrying only superscripts or only subscripts are referred to as the contravariant<strong>and</strong> covariant components respectively; all others are called mixed components.26.15 The metric tensorAny particular curvilinear coordinate system is completely characterised at eachpoint in space by the nine quantitiesg ij = e i · e j , (26.56)which, as we will show, are the covariant components of a symmetric second-ordertensor g called the metric tensor.Since an infinitesimal vector displacement can be written as dr = du i e i , we findthat the square of the infinitesimal arc length (ds) 2 can be written in terms of themetric tensor as(ds) 2 = dr · dr = du i e i · du j e j = g ij du i du j . (26.57)It may further be shown that the volume element dV is given bydV = √ gdu 1 du 2 du 3 , (26.58)where g is the determinant of the matrix [ g ij ], which has the covariant componentsof the metric tensor as its elements.If we compare equations (26.57) <strong>and</strong> (26.58) with the analogous ones in section10.10 then we see that in the special case where the coordinate system is orthogonal(so that e i · e j =0<strong>for</strong>i ≠ j) the metric tensor can be written in terms of thecoordinate-system scale factors h i , i =1, 2, 3as{h2g ij = i i = j,0 i ≠ j.Its determinant is then given by g = h 2 1 h2 2 h2 3 .957


TENSORS◮Calculate the elements g ij of the metric tensor <strong>for</strong> cylindrical polar coordinates. Hencefind the square of the infinitesimal arc length (ds) 2 <strong>and</strong> the volume dV <strong>for</strong> this coordinatesystem.As discussed in section 10.9, in cylindrical polar coordinates (u 1 ,u 2 ,u 3 )=(ρ, φ, z) <strong>and</strong>sothe position vector r of any point P may be writtenr = ρ cos φ i + ρ sin φ j + z k.From this we obtain the (covariant) basis vectors:e 1 = ∂r = cosφ i +sinφ j;∂ρe 2 = ∂r = −ρ sin φ i + ρ cos φ j;∂φe 3 = ∂r = k. (26.59)∂zThus the components of the metric tensor [g ij ]=[e i · e j ] are found to be⎛G =[g ij ]= ⎝ 1 0 0⎞0 ρ 2 0 ⎠ , (26.60)0 0 1from which we see that, as expected <strong>for</strong> an orthogonal coordinate system, the metric tensoris diagonal, the diagonal elements being equal to the squares of the scale factors of thecoordinate system.From (26.57), the square of the infinitesimal arc length in this coordinate system is givenby(ds) 2 = g ij du i du j =(dρ) 2 + ρ 2 (dφ) 2 +(dz) 2 ,<strong>and</strong>, using (26.58), the volume element is found to bedV = √ gdu 1 du 2 du 3 = ρdρdφdz.These expressions are identical to those derived in section 10.9. ◭We may also express the scalar product of two vectors in terms of the metrictensor:a · b = a i e i · b j e j = g ij a i b j , (26.61)where we have used the contravariant components of the two vectors. Similarly,using the covariant components, we can write the same scalar product asa · b = a i e i · b j e j = g ij a i b j , (26.62)where we have defined the nine quantities g ij = e i ·e j . As we shall show, they <strong>for</strong>mthe contravariant components of the metric tensor g <strong>and</strong> are, in general, differentfrom the quantities g ij . Finally, we could express the scalar product in terms ofthe contravariant components of one vector <strong>and</strong> the covariant components of theother,a · b = a i e i · b j e j = a i b j δ i j = a i b i , (26.63)958


26.15 THE METRIC TENSORwhere we have used the reciprocity relation (26.54). Similarly, we could writea · b = a i e i · b j e j = a i b j δ j i = a i b i . (26.64)By comparing the four alternative expressions (26.61)–(26.64) <strong>for</strong> the scalarproduct of two vectors we can deduce one of the most useful properties ofthe quantities g ij <strong>and</strong> g ij . Since g ij a i b j = a i b i holds <strong>for</strong> any arbitrary vectorcomponents a i , it follows thatg ij b j = b i ,which illustrates the fact that the covariant components g ij ofthemetrictensorcan be used to lower an index. In other words, it provides a means of obtainingthe covariant components of a vector from its contravariant components. By asimilar argument, we haveg ij b j = b i ,so that the contravariant components g ij can be used to per<strong>for</strong>m the reverseoperation of raising an index.It is straight<strong>for</strong>ward to show that the contravariant <strong>and</strong> covariant basis vectors,e i <strong>and</strong> e i respectively, are related in the same way as other vectors, i.e. bye i = g ij e j <strong>and</strong> e i = g ij e j .We also note that, since e i <strong>and</strong> e i are reciprocal systems of vectors in threedimensionalspace (see chapter 7), we may writee i =e j × e ke i · (e j × e k ) ,<strong>for</strong> the combination of subscripts i, j, k =1, 2, 3 <strong>and</strong> its cyclic permutations. Asimilar expression holds <strong>for</strong> e i in terms of the e i -basis. Moreover, it may be shownthat |e 1 · (e 2 × e 3 )| = √ g.◮Show that the matrix [g ij ] is the inverse of the matrix [g ij ]. Hence calculate the contravariantcomponents g ij of the metric tensor in cylindrical polar coordinates.Using the index-lowering <strong>and</strong> index-raising properties of g ij <strong>and</strong> g ij on an arbitrary vectora, we findδka i k = a i = g ij a j = g ij g jk a k .But, since a is arbitrary, we must haveg ij g jk = δk. i (26.65)Denoting the matrix [g ij ]byG <strong>and</strong> [g ij ]byĜ, equation (26.65) can be written in matrix<strong>for</strong>m as ĜG = I, whereI is the unit matrix. Hence G <strong>and</strong> Ĝ are inverse matrices of eachother.959


TENSORSThus, by inverting the matrix G in (26.60), we find that the elements g ij are given incylindrical polar coordinates byĜ =[g ij ]=⎛⎝ 1 0 0 1/ρ 00⎞⎠ . ◭0 0 1So far we have not considered the components of the metric tensor gj i with onesubscript <strong>and</strong> one superscript. By analogy with (26.56), these mixed componentsare given bygj i = e i · e j = δ j i ,<strong>and</strong> so the components of gji are identical to those of δj i.Wemaythere<strong>for</strong>econsider the δj i to be the mixed components of the metric tensor g.26.16 General coordinate trans<strong>for</strong>mations <strong>and</strong> tensorsWe now discuss the concept of general trans<strong>for</strong>mations from one coordinatesystem, u 1 ,u 2 ,u 3 , to another, u ′1 ,u ′2 ,u ′3 . We can describe the coordinate trans<strong>for</strong>musing the three equationsu ′i = u ′i (u 1 ,u 2 ,u 3 ),<strong>for</strong> i =1, 2, 3, in which the new coordinates u ′i can be arbitrary functions of the oldones u i rather than just represent linear orthogonal trans<strong>for</strong>mations (rotations)of the coordinate axes. We shall assume also that the trans<strong>for</strong>mation can beinverted, so that we can write the old coordinates in terms of the new ones asu i = u i (u ′1 ,u ′2 ,u ′3 ),As an example, we may consider the trans<strong>for</strong>mation from spherical polar toCartesian coordinates, given byx = r sin θ cos φ,y = r sin θ sin φ,z = r cos θ,which is clearly not a linear trans<strong>for</strong>mation.The two sets of basis vectors in the new coordinate system, u ′1 ,u ′2 ,u ′3 , are givenas in (26.55) bye ′ i = ∂r <strong>and</strong> e ′i = ∇u ′i . (26.66)∂u ′iConsidering the first set, we have from the chain rule that∂r∂u j = ∂u′ i∂r∂u j ∂u ′ i ,960


26.16 GENERAL COORDINATE TRANSFORMATIONS AND TENSORSso that the basis vectors in the old <strong>and</strong> new coordinate systems are related bye j = ∂u′ i∂u j e′ i. (26.67)Now, since we can write any arbitrary vector a in terms of either basis asa = a ′i e ′ i = a j e j = a j ∂u′i∂u j e′ i,it follows that the contravariant components of a vector must trans<strong>for</strong>m asa ′i = ∂u′ i∂u j a j . (26.68)In fact, we use this relation as the defining property <strong>for</strong> a set of quantities a i to<strong>for</strong>m the contravariant components of a vector.◮Find an expression analogous to (26.67) relating the basis vectors e i <strong>and</strong> e ′i in the twocoordinate systems. Hence deduce the way in which the covariant components of a vectorchange under a coordinate trans<strong>for</strong>mation.If we consider the second set of basis vectors in (26.66), e ′i = ∇u ′i , we have from the chainrule that∂u j∂x = ∂u j ∂u ′i∂u ′i ∂x<strong>and</strong> similarly <strong>for</strong> ∂u j /∂y <strong>and</strong> ∂u j /∂z. So the basis vectors in the old <strong>and</strong> new coordinatesystems are related bye j = ∂u j∂u ′i e′i . (26.69)For any arbitrary vector a,a = a ′ ie ′i = a j e j = a j∂u je′i∂u′i<strong>and</strong> so the covariant components of a vector must trans<strong>for</strong>m asa ′ i = ∂u j∂u a j. (26.70)′iAnalogously to the contravariant case (26.68), we take this result as the defining propertyof the covariant components of a vector. ◭We may compare the trans<strong>for</strong>mation laws (26.68) <strong>and</strong> (26.70) with those <strong>for</strong>a first-order Cartesian tensor under a rigid rotation of axes. Let us considera rotation of Cartesian axes x i throughanangleθ about the 3-axis to a newset x ′i , i =1, 2, 3, as given by (26.7) <strong>and</strong> the inverse trans<strong>for</strong>mation (26.8). It isstraight<strong>for</strong>ward to show that∂x j= ∂x′ i∂x ′ i∂x j = L ij,961


TENSORSwhere the elements L ij are given by⎛L = ⎝cos θ sin θ⎞0− sin θ cos θ 0 ⎠ .0 0 1Thus (26.68) <strong>and</strong> (26.70) agree with our earlier definition in the special case of arigid rotation of Cartesian axes.Following on from (26.68) <strong>and</strong> (26.70), we proceed in a similar way to definegeneral tensors of higher rank. For example, the contravariant, mixed <strong>and</strong>covariant components, respectively, of a second-order tensor must trans<strong>for</strong>m asfollows:contravariant components,mixed components,covariant components,T ′ ij = ∂u′ i∂u k ∂u ′ j∂u l T kl ;T ′ ij = ∂u′ i∂u k ∂u l∂u ′ j T k l ;T ′ ij = ∂uk∂u ′ i∂u l∂u ′ j T kl.It is important to remember that these quantities <strong>for</strong>m the components of thesame tensor T but refer to different tensor bases made up from the basis vectorsof the different coordinate systems. For example, in terms of the contravariantcomponents we may writeT = T ij e i ⊗ e j = T ′ ij e ′ i ⊗ e ′ j.We can clearly go on to define tensors of higher order, with arbitrary numbersof covariant (subscript) <strong>and</strong> contravariant (superscript) indices, by dem<strong>and</strong>ingthat their components trans<strong>for</strong>m as follows:T ′ ij···klm···n = ∂u′ i∂u ′ j···∂u′ k∂u d ∂u e ∂uf∂u a ∂u b ∂u c ···∂u ′ l∂u′m∂u ′n T ab···c de···f. (26.71)Using the revised summation convention described in section 26.14, the algebraof general tensors is completely analogous to that of the Cartesian tensorsdiscussed earlier. For example, as with Cartesian coordinates, the Kroneckerdelta is a tensor provided it is written as the mixed tensor δj i sinceδ ′i j = ∂u′i ∂u l∂u k ∂u ′j δk l = ∂u′i ∂u k ∂u′i∂u k =∂u′j∂u = ′j δi j,where we have used the chain rule to justify the third equality. This also showsthat δj i is isotropic. As discussed at the end of section 26.15, the δj i can beconsidered as the mixed components of the metric tensor g.962


26.17 RELATIVE TENSORS◮Show that the quantities g ij = e i · e jtensor.<strong>for</strong>m the covariant components of a second-orderIn the new (primed) coordinate system we haveg ij ′ = e ′ i · e ′ j,but using (26.67) <strong>for</strong> the inverse trans<strong>for</strong>mation, we havee ′ i = ∂uk∂u e k,′i<strong>and</strong> similarly <strong>for</strong> e ′ j .Thuswemaywriteg ij ′ = ∂uk ∂u le∂u ′i ∂u ′j k · e l = ∂uk ∂u l∂u ′i ∂u g kl,′jwhich shows that the g ij are indeed the covariant components of a second-order tensor(themetrictensorg). ◭A similar argument to that used in the above example shows that the quantitiesg ij <strong>for</strong>m the contravariant components of a second-order tensor which trans<strong>for</strong>msaccording tog ′ ij = ∂u′ i∂u ′ j∂u k ∂u l gkl .In the previous section we discussed the use of the components g ij <strong>and</strong> g ij inthe raising <strong>and</strong> lowering of indices in contravariant <strong>and</strong> covariant vectors. Thiscan be extended to tensors of arbitrary rank. In general, contraction of a tensorwith g ij will convert the contracted index from being contravariant (superscript)to covariant (subscript), i.e. it is lowered. This can be repeated <strong>for</strong> as many indicesare required. For example,Similarly contraction with g ij raises an index, i.e.T ij = g ik T k j = g ik g jl T kl . (26.72)T ij = g ik T jk = gik g jl T kl . (26.73)That (26.72) <strong>and</strong> (26.73) are mutually consistent may be shown by using the factthat g ik g kj = δ i j .26.17 Relative tensorsIn section 26.10 we introduced the concept of pseudotensors in the context of therotation (proper or improper) of a set of Cartesian axes. Generalising to arbitrarycoordinate trans<strong>for</strong>mations leads to the notion of a relative tensor.For an arbitrary coordinate trans<strong>for</strong>mation from one general coordinate system963


TENSORSu i to another u ′i , we may define the Jacobian of the trans<strong>for</strong>mation (see chapter 6)as the determinant of the trans<strong>for</strong>mation matrix [∂u ′i /∂u j ]: this is usually denotedbyJ =∂u ′∣ ∂u ∣ .Alternatively, we may interchange the primed <strong>and</strong> unprimed coordinates toobtain |∂u/∂u ′ | =1/J: un<strong>for</strong>tunately this also is often called the Jacobian of thetrans<strong>for</strong>mation.Using the Jacobian J, we define a relative tensor of weight w as one whosecomponents trans<strong>for</strong>m as follows:T ′ ij···klm···n = ∂u′ i∂u ′ j···∂u′ k∣∂u d ∂u e ∂uf∂u a ∂u b ∂u c ∂u ′ l∂u ′ m ···∂u ′ n T ab···c ∂u ∣∣∣wde···f ∣∂u ′ .(26.74)Comparing this expression with (26.71), we see that a true (or absolute) generaltensor may be considered as a relative tensor of weight w =0.Ifw = −1, on theother h<strong>and</strong>, the relative tensor is known as a general pseudotensor <strong>and</strong> if w =1as a tensor density.It is worth comparing (26.74) with the definition (26.39) of a Cartesian pseudotensor.For the latter, we are concerned only with its behaviour under a rotation(proper or improper) of Cartesian axes, <strong>for</strong> which the Jacobian J = ±1. Thus,general relative tensors of weight w = −1 <strong>and</strong>w = 1 would both satisfy thedefinition (26.39) of a Cartesian pseudotensor.◮If the g ij are the covariant components of the metric tensor, show that the determinant gof the matrix [g ij ] is a relative scalar of weight w =2.The components g ij trans<strong>for</strong>m asg ij ′ = ∂uk ∂u l∂u ′i ∂u g kl.′jDefining the matrices U =[∂u i /∂u ′j ], G =[g ij ]<strong>and</strong>G ′ =[g ij ′ ], we may write this expressionasG ′ = U T GU.Taking the determinant of both sides, we obtain∣ g ′ = |U| 2 g =∂u ∣∣∣2∣ g,∂u ′which shows that g is a relative scalar of weight w =2.◭From the discussion in section 26.8, it can be seen that ɛ ijk is a covariantrelative tensor of weight −1. We may also define the contravariant tensor ɛ ijk ,which is numerically equal to ɛ ijk but is a relative tensor of weight +1.If two relative tensors have weights w 1 <strong>and</strong> w 2 respectively then, from (26.74),964


26.18 DERIVATIVES OF BASIS VECTORS AND CHRISTOFFEL SYMBOLSthe outer product of the two tensors, or any contraction of them, is a relativetensor of weight w 1 + w 2 . As a special case, we may use ɛ ijk <strong>and</strong> ɛ ijk to constructpseudovectors from antisymmetric tensors <strong>and</strong> vice versa, in an analogous wayto that discussed in section 26.11.For example, if the A ij are the contravariant components of an antisymmetrictensor (w =0)thenp i = 1 2 ɛ ijkA jkare the covariant components of a pseudovector (w = −1), since ɛ ijk has weightw = −1. Similarly, we may show thatA ij = ɛ ijk p k .26.18 Derivatives of basis vectors <strong>and</strong> Christoffel symbolsIn Cartesian coordinates, the basis vectors e i are constant <strong>and</strong> so their derivativeswith respect to the coordinates vanish. In a general coordinate system, however,the basis vectors e i <strong>and</strong> e i are functions of the coordinates. There<strong>for</strong>e, in orderthat we may differentiate general tensors we must consider the derivatives of thebasis vectors.First consider the derivative ∂e i /∂u j . Since this is itself a vector, it can bewritten as a linear combination of the basis vectors e k , k =1, 2, 3. If we introducethe symbol Γ k ijto denote the coefficients in this combination, we have∂e i∂u j =Γk ije k . (26.75)The coefficient Γ k ij is the kth component of the vector ∂e i/∂u j .Usingthereciprocityrelation e i · e j = δj i , these 27 numbers are given (at each point in space)byΓ k ij = e k · ∂e i∂u j . (26.76)Furthermore, by differentiating the reciprocity relation e i · e j = δj i with respectto the coordinates, <strong>and</strong> using (26.76), it is straight<strong>for</strong>ward to show that thederivatives of the contravariant basis vectors are given by∂e i∂u j = −Γi kje k . (26.77)The symbol Γ k ijis called a Christoffel symbol (of the second kind), but, despiteappearances to the contrary, these quantities do not <strong>for</strong>m the components of athird-order tensor. It is clear from (26.76) that in Cartesian coordinates Γ k ij =0<strong>for</strong> all values of the indices i, j <strong>and</strong> k.965


TENSORS◮Using (26.76), deduce the way in which the quantities Γ k ijtrans<strong>for</strong>m under a generalcoordinate trans<strong>for</strong>mation, <strong>and</strong> hence show that they do not <strong>for</strong>m the components of athird-order tensor.In a new coordinate systemΓ ′k ij = e ′k · ∂e′ i∂u , ′jbut from (26.69) <strong>and</strong> (26.67) respectively we have, on reversing primed <strong>and</strong> unprimedvariables,e ′k = ∂u′k∂u n en <strong>and</strong> e ′ i = ∂ul∂u e l.′iThere<strong>for</strong>e in the new coordinate system the quantities Γ ′k ij are given byΓ ′k ij = ∂u′k∂u n en ·= ∂u′k∂u n en ·( )∂ ∂ul∂u ′j ∂u e ′i l( )∂ 2 u l∂u ′j ∂u e ′i l + ∂ul ∂e l∂u ′i ∂u ′j= ∂u′k ∂ 2 u le n · e∂u n ∂u ′j ∂u ′i l + ∂u′k ∂u l ∂u me n · ∂e l∂u n ∂u ′i ∂u ′j ∂u m= ∂u′k ∂ 2 u l+ ∂u′k ∂u l ∂u m∂u l ∂u ′j ∂u ′i ∂u n ∂u ′i ∂u ′j Γn lm, (26.78)where in the last line we have used (26.76) <strong>and</strong> the reciprocity relation e n · e l = δl n.From(26.78), because of the presence of the first term on the right-h<strong>and</strong> side, we concludeimmediately that the Γ k ijdo not <strong>for</strong>m the components of a third-order tensor. ◭In a given coordinate system, in principle we may calculate the Γ k ijusing(26.76). In practice, however, it is often quicker to use an alternative expression,which we now derive, <strong>for</strong> the Christoffel symbol in terms of the metric tensor g ij<strong>and</strong> its derivatives with respect to the coordinates.Firstly we note that the Christoffel symbol Γ k ijis symmetric with respect tothe interchange of its two subscripts i <strong>and</strong> j. This is easily shown: since∂e i∂u j =∂2 r∂u j ∂u i =∂2 r∂u i ∂u j = ∂e j∂u i ,it follows from (26.75) that Γ k ij e k =Γ k ji e k. Taking the scalar product with e l <strong>and</strong>using the reciprocity relation e k · e l = δk l gives immediately thatΓ l ij =Γ l ji.To obtain an expression <strong>for</strong> Γ k ijwe then use g ij = e i · e j <strong>and</strong> consider thederivative∂g ij∂u k = ∂e i∂u k · e j + e i · ∂e j∂u k=Γ l ike l · e j + e i · Γ l jke l=Γ l ikg lj +Γ l jkg il , (26.79)966


26.18 DERIVATIVES OF BASIS VECTORS AND CHRISTOFFEL SYMBOLSwhere we have used the definition (26.75). By cyclically permuting the free indicesi, j, k in (26.79), we obtain two further equivalent relations,<strong>and</strong>∂g jk∂u i =Γ l jig lk +Γ l kig jl (26.80)∂g ki∂u j =Γ l kjg li +Γ l ijg kl . (26.81)If we now add (26.80) <strong>and</strong> (26.81) together <strong>and</strong> subtract (26.79) from the result,we find∂g jk∂u i+ ∂g ki∂u j − ∂g ij∂u k =Γl jig lk +Γ l kig jl +Γ l kjg li +Γ l ijg kl − Γ l ikg lj − Γ l jkg il=2Γ l ijg kl ,where we have used the symmetry properties of both Γ l ij <strong>and</strong> g ij. Contractingboth sides with g mk leads to the required expression <strong>for</strong> the Christoffel symbol interms of the metric tensor <strong>and</strong> its derivatives, namelyΓ m ij = 1 2 gmk (∂gjk∂u i+ ∂g ki∂u j − ∂g )ij∂u k . (26.82)◮Calculate the Christoffel symbols Γ m ij <strong>for</strong> cylindrical polar coordinates.We may use either (26.75) or (26.82) to calculate the Γ m ij <strong>for</strong> this simple coordinate system.In cylindrical polar coordinates (u 1 ,u 2 ,u 3 )=(ρ, φ, z), the basis vectors e i are given by(26.59). It is straight<strong>for</strong>ward to show that the only derivatives of these vectors with respectto the coordinates that are non-zero are∂e ρ∂φ = 1 ρ e φ,∂e φ∂ρ = 1 ρ e φ,∂e φ∂φ = −ρe ρ.Thus, from (26.75), we have immediately thatΓ 2 12 =Γ 2 21 = 1 ρ<strong>and</strong> Γ 1 22 = −ρ. (26.83)Alternatively, using (26.82) <strong>and</strong> the fact that g 11 =1,g 22 = ρ 2 , g 33 = 1 <strong>and</strong> the othercomponents are zero, we see that the only three non-zero Christoffel symbols are indeedΓ 2 12 =Γ2 21 <strong>and</strong> Γ1 22 .ThesearegivenbyΓ 2 12 =Γ 2 21 = 1 ∂g 222g 22 ∂u = 1 ∂1 2ρ 2 ∂ρ (ρ2 )= 1 ρ ,Γ 1 22 = − 1 ∂g 222g 11 ∂u = − 1 ∂1 2 ∂ρ (ρ2 )=−ρ,which agree with the expressions found directly from (26.75) <strong>and</strong> given in (26.83 ). ◭967


TENSORS26.19 Covariant differentiationFor Cartesian tensors we noted that the derivative of a scalar is a (covariant)vector.Thisisalsotrue<strong>for</strong>general tensors, as may be shown by considering thedifferential of a scalardφ = ∂φ∂u i dui .Since the du i are the components of a contravariant vector <strong>and</strong> dφ is a scalar,we have by the quotient law, discussed in section 26.7, that the quantities ∂φ/∂u imust <strong>for</strong>m the components of a covariant vector. As a second example, if thecontravariant components in Cartesian coordinates of a vector v are v i , then thequantities ∂v i /∂x j <strong>for</strong>m the components of a second-order tensor.However, it is straight<strong>for</strong>ward to show that in non-Cartesian coordinates differentiationof the components of a general tensor, other than a scalar, with respectto the coordinates does not in general result in the components of another tensor.◮Show that, in general coordinates, the quantities ∂v i /∂u j do not <strong>for</strong>m the components ofa tensor.We may show this directly by considering( ) ∂vi ′= ∂v′i= ∂uk ∂v ′i∂u j ∂u ′j ∂u ′j ∂u k( )= ∂uk ∂ ∂u ′i∂u ′j ∂u k ∂u l vl= ∂uk ∂u ′i ∂v l∂u ′j ∂u l ∂u + ∂uk ∂ 2 u ′ik ∂u ′j ∂u k ∂u l vl . (26.84)The presence of the second term on the right-h<strong>and</strong> side of (26.84) shows that the ∂v i /∂x jdo not <strong>for</strong>m the components of a second-order tensor. This term arises because the‘trans<strong>for</strong>mation matrix’ [∂u ′i /∂u j ] changes as the position in space at which it is evaluatedis changed. This is not true in Cartesian coordinates, <strong>for</strong> which the second term vanishes<strong>and</strong> ∂v i /∂x j is a second-order tensor. ◭We may, however, use the Christoffel symbols discussed in the previous sectionto define a new covariant derivative of the components of a tensor that doesresult in the components of another tensor.Let us first consider the derivative of a vector v with respect to the coordinates.Writing the vector in terms of its contravariant components v = v i e i , we find∂v∂u j = ∂vi∂u j e i + v i ∂e i∂u j , (26.85)where the second term arises because, in general, the basis vectors e i are not968


26.19 COVARIANT DIFFERENTIATIONconstant (this term vanishes in Cartesian coordinates). Using (26.75) we write∂v∂u j = ∂vi∂u j e i + v i Γ k ije k .Since i <strong>and</strong> k are dummy indices in the last term on the right-h<strong>and</strong> side, we mayinterchange them to obtain( )∂v∂u j = ∂vi∂v∂u j e i + v k Γ i ikje i =∂u j + vk Γ i kj e i . (26.86)The reason <strong>for</strong> the interchanging the dummy indices, as shown in (26.86), is thatwe may now factor out e i . The quantity in parentheses is called the covariantderivative, <strong>for</strong> which the st<strong>and</strong>ard notation isv i ; j ≡ ∂vi∂u j +Γi kjv k , (26.87)the semicolon subscript denoting covariant differentiation. A similar short-h<strong>and</strong>notation also exists <strong>for</strong> the partial derivatives, a comma being used <strong>for</strong> theseinstead of a semicolon; <strong>for</strong> example, ∂v i /∂u j is denoted by v i ,j. In Cartesiancoordinates all the Γ i kjare zero, <strong>and</strong> so the covariant derivative reduces to thesimple partial derivative ∂v i /∂u j .Using the short-h<strong>and</strong> semicolon notation, the derivative of a vector may bewritten in the very compact <strong>for</strong>m∂v∂u j = vi ; je i<strong>and</strong>, by the quotient rule (section 26.7), it is clear that the v i ; j are the (mixed)components of a second-order tensor. This may also be verified directly, usingthe trans<strong>for</strong>mation properties of ∂v i /∂u j <strong>and</strong> Γ i kjgiven in (26.84) <strong>and</strong> (26.78)respectively.In general, we may regard the v i ; j as the mixed components of a secondordertensor called the covariant derivative of v <strong>and</strong> denoted by ∇v. In Cartesiancoordinates, the components of this tensor are just ∂v i /∂x j .◮Calculate v i ; i in cylindrical polar coordinates.Contracting (26.87) we obtainNow from (26.83) we havev i ; i = ∂vi∂u i +Γ i kiv k .Γ i 1i =Γ 1 11 +Γ 2 12 +Γ 3 13 =1/ρ,Γ i 2i =Γ 1 21 +Γ 2 22 +Γ 3 23 =0,Γ i 3i =Γ 1 31 +Γ 2 32 +Γ 3 33 =0,969


TENSORS<strong>and</strong> sov i ; i = ∂vρ∂ρ + ∂vφ∂φ + ∂vz∂z + 1 ρ vρ= 1 ∂ρ ∂ρ (ρvρ )+ ∂vφ∂φ + ∂vz∂z .This result is identical to the expression <strong>for</strong> the divergence of a vector field in cylindricalpolar coordinates given in section 10.9. This is discussed further in section 26.20. ◭So far we have considered only the covariant derivative of the contravariantcomponents v i of a vector. The corresponding result <strong>for</strong> the covariant componentsv i may be found in a similar way, by considering the derivative of v = v i e i <strong>and</strong>using (26.77) to obtainv i; j = ∂v i∂u j − Γk ijv k . (26.88)Comparing the expressions (26.87) <strong>and</strong> (26.88) <strong>for</strong> the covariant derivative ofthe contravariant <strong>and</strong> covariant components of a vector respectively, we see thatthere are some similarities <strong>and</strong> some differences. It may help to remember thatthe index with respect to which the covariant derivative is taken (j in this case), isalso the last subscript on the Christoffel symbol; the remaining indices can thenbe arranged in only one way without raising or lowering them. It only remains tonote that <strong>for</strong> a covariant index (subscript) the Christoffel symbol carries a minussign, whereas <strong>for</strong> a contravariant index (superscript) the sign is positive.Following a similar procedure to that which led to equation (26.87), we mayobtain expressions <strong>for</strong> the covariant derivatives of higher-order tensors.◮By considering the derivative of the second-order tensor T with respect to the coordinateu k , find an expression <strong>for</strong> the covariant derivative T ij ; kof its contravariant components.Expressing T in terms of its contravariant components, we have∂T∂u = ∂k ∂u (T ij e k i ⊗ e j )ij∂T=∂u e k i ⊗ e j + T ij ∂e i∂u ⊗ e k j + T ij e i ⊗ ∂e j∂u . kUsing (26.75), we can rewrite the derivatives of the basis vectors in terms of Christoffelsymbols to obtainij∂T ∂T=∂uk ∂u e k i ⊗ e j + T ij Γ l ike l ⊗ e j + T ij e i ⊗ Γ l jke l .Interchanging the dummy indices i <strong>and</strong> l in the second term <strong>and</strong> j <strong>and</strong> l in the third termon the right-h<strong>and</strong> side, this becomes( )∂T ∂Tij∂u = +Γ i k ∂ulkT lj +Γ j k lk T il e i ⊗ e j ,970


26.20 VECTOR OPERATORS IN TENSOR FORMwhere the expression in parentheses is the required covariant derivativeT ij ij∂T; k= +Γ i∂ulkT lj +Γ j k lk T il . (26.89)Using (26.89), the derivative of the tensor T with respect to u k cannowbewrittenintermsof its contravariant components as∂T∂u = T ij k ; k e i ⊗ e j . ◭Results similar to (26.89) may be obtained <strong>for</strong> the the covariant derivatives ofthe mixed <strong>and</strong> covariant components of a second-order tensor. Collecting theseresults together, we haveT ij ; k = T ij ,k +Γi lkT lj +Γ j lk T il ,T i j; k = T i j,k +Γ i lkT l j − Γ l jkT i l,T ij; k = T ij, k − Γ l ikT lj − Γ l jkT il ,where we have used the comma notation <strong>for</strong> partial derivatives. The position ofthe indices in these expressions is very systematic: <strong>for</strong> each contravariant index(superscript) on the LHS we add a term on the RHS containing a Christoffelsymbol with a plus sign, <strong>and</strong> <strong>for</strong> every covariant index (subscript) we add acorresponding term with a minus sign. This is extended straight<strong>for</strong>wardly totensors with an arbitrary number of contravariant <strong>and</strong> covariant indices.We note that the quantities T ij ; k , T i j; k <strong>and</strong> T ij; k are the components of thesame third-order tensor ∇T with respect to different tensor bases, i.e.∇T = T ij ; k e i ⊗ e j ⊗ e k = T i j; ke i ⊗ e j ⊗ e k = T ij; k e i ⊗ e j ⊗ e k .We conclude this section by considering briefly the covariant derivative of ascalar. The covariant derivative differs from the simple partial derivative withrespect to the coordinates only because the basis vectors of the coordinatesystem change with position in space (hence <strong>for</strong> Cartesian coordinates there is nodifference). However, a scalar φ does not depend on the basis vectors at all <strong>and</strong>so its covariant derivative must be the same as its partial derivative, i.e.φ ; j = ∂φ∂u j = φ ,j. (26.90)26.20 Vector operators in tensor <strong>for</strong>mIn section 10.10 we used vector calculus methods to find expressions <strong>for</strong> vectordifferential operators, such as grad, div, curl <strong>and</strong> the Laplacian, in general orthogonalcurvilinear coordinates, taking cylindrical <strong>and</strong> spherical polars as particularexamples. In this section we use the framework of general tensors that we havedeveloped to obtain, in tensor <strong>for</strong>m, expressions <strong>for</strong> these operators that are validin all coordinate systems, whether orthogonal or not.971


TENSORSIn order to compare the results obtained here with those given in section10.10 <strong>for</strong> orthogonal coordinates, it is necessary to remember that here we areworking with the (in general) non-unit basis vectors e i = ∂r/∂u i or e i = ∇u i .Thus the components of a vector v = v i e i are not the same as the components ˆv iappropriate to the corresponding unit basis ê i . In fact, if the scale factors of thecoordinate system are h i , i =1, 2, 3, then v i = ˆv i /h i (no summation over i).As mentioned in section 26.15, <strong>for</strong> an orthogonal coordinate system with scalefactors h i we have{h2g ij = i if i = j,0 otherwise{1/h<strong>and</strong> g ij 2= i if i = j,0 otherwise,<strong>and</strong> so the determinant g of the matrix [g ij ] is given by g = h 2 1 h2 2 h2 3 .GradientThe gradient of a scalar φ is given by∇φ = φ ; i e i = ∂φ∂u i ei , (26.91)since the covariant derivative of a scalar is the same as its partial derivative.DivergenceReplacing the partial derivatives that occur in Cartesian coordinates with covariantderivatives, the divergence of a vector field v in a general coordinate systemis given by∇ · v = v i ; i = ∂vi∂u i +Γ i kiv k .Using the expression (26.82) <strong>for</strong> the Christoffel symbol in terms of the metrictensor, we find(Γ i ki = 1 ∂gil2 gil ∂u k + ∂g kl∂u i − ∂g )ki∂u l = 1 ∂g 2 gil il∂u k . (26.92)The last two terms have cancelled becauseg il ∂g kl∂u i = g li ∂g ki∂u l = g il ∂g ki∂u l ,where in the first equality we have interchanged the dummy indices i <strong>and</strong> l, <strong>and</strong>in the second equality have used the symmetry of the metric tensor.We may simplify (26.92) still further by using a result concerning the derivativeof the determinant of a matrix whose elements are functions of the coordinates.972


26.20 VECTOR OPERATORS IN TENSOR FORM◮Suppose A =[a ij ], B =[b ij ] <strong>and</strong> that B = A −1 . By considering the determinant a = |A|,show that∂a∂u = ab ji ∂a ijk ∂u . kIf we denote the cofactor of the element a ij by ∆ ij then the elements of the inverse matrixare given by (see chapter 8)However, the determinant of A is given byb ij = 1 a ∆ji . (26.93)a = ∑ ja ij ∆ ij ,in which we have fixed i <strong>and</strong> written the sum over j explicitly, <strong>for</strong> clarity. Partiallydifferentiating both sides with respect to a ij , we then obtain∂a=∆ ij , (26.94)∂a ijsince a ij does not occur in any of the cofactors ∆ ij .Now, if the a ij depend on the coordinates then so will the determinant a <strong>and</strong>, by thechain rule, we have∂a∂u = ∂a ∂a ij ∂ak ∂a ij ∂u k =∆ij ij∂u = ab ji ∂a ijk ∂u , (26.95)kin which we have used (26.93) <strong>and</strong> (26.94). ◭Applying the result (26.95) to the determinant g of the metric tensor, <strong>and</strong>remembering both that g ik g kj = δ i j <strong>and</strong> that gij is symmetric, we obtain∂g∂u k = ∂g ggij ij∂u k . (26.96)Substituting (26.96) into (26.92) we find that the expression <strong>for</strong> the Christoffelsymbol can be much simplified to giveΓ i ki = 1 ∂g2g ∂u k = √ 1 ∂ √ gg ∂u k .Thus finally we obtain the expression <strong>for</strong> the divergence of a vector field in ageneral coordinate system as∇ · v = v i ; i = √ 1 ∂g ∂u j (√ gv j ). (26.97)LaplacianIf we replace v by ∇φ in ∇ · v then we obtain the Laplacian ∇ 2 φ. From (26.91),we havev i e i = v = ∇φ = ∂φ∂u i ei ,973


TENSORS<strong>and</strong> so the covariant components of v are given by v i = ∂φ/∂u i . In (26.97),however, we require the contravariant components v i . These may be obtained byraising the index using the metric tensor, to givev j = g jk jk ∂φv k = g∂u k .Substituting this into (26.97) we obtain∇ 2 φ = √ 1 (∂ √gg jk ∂φ )g ∂u j ∂u k . (26.98)◮Use (26.98) to find the expression <strong>for</strong> ∇ 2 φ in an orthogonal coordinate system with scalefactors h i , i =1, 2, 3.For an orthogonal coordinate system √ g = h 1 h 2 h 3 ;further,g ij =1/h 2 i if i = j <strong>and</strong> g ij =0otherwise. There<strong>for</strong>e, from (26.98) we have( )∇ 2 φ = 1 ∂ h 1 h 2 h 3 ∂φ,h 1 h 2 h 3 ∂u j h 2 j∂u jwhich agrees with the results of section 10.10. ◭CurlThe special vector <strong>for</strong>m of the curl of a vector field exists only in three dimensions.We there<strong>for</strong>e consider a more general <strong>for</strong>m valid in higher-dimensional spaces aswell. In a general space the operation curl v is defined by(curl v) ij = v i; j − v j; i ,which is an antisymmetric covariant tensor.In fact the difference of derivatives can be simplified, sincev i; j − v j; i = ∂v i∂u j − Γl ijv l − ∂v j∂u i +Γ l jiv l= ∂v i∂u j − ∂v j∂u i ,where the Christoffel symbols have cancelled because of their symmetry properties.Thus curl v can be written in terms of partial derivatives as(curl v) ij = ∂v i∂u j − ∂v j∂u i .Generalising slightly the discussion of section 26.17, in three dimensions we mayassociate with this antisymmetric second-order tensor a vector with contravariantcomponents,(∇×v) i = − 12 √ g ɛijk (curl v) jk= − 1 (2 √ ∂vjg ɛijk ∂u k − ∂v )k∂u j = √ 1 ɛ ijk ∂v kg ∂u j ;974


26.21 ABSOLUTE DERIVATIVES ALONG CURVESthis is the analogue of the expression in Cartesian coordinates discussed insection 26.8.26.21 Absolute derivatives along curvesIn section 26.19 we discussed how to differentiate a general tensor with respectto the coordinates <strong>and</strong> introduced the covariant derivative. In this section weconsider the slightly different problem of calculating the derivative of a tensoralong a curve r(t) that is parameterised by some variable t.Let us begin by considering the derivative of a vector v along the curve. If weintroduce an arbitrary coordinate system u i with basis vectors e i , i =1, 2, 3, thenwe may write v = v i e i <strong>and</strong> so obtaindvdt = dvidt e i + v i de idt= dvidt e i + v i ∂e i du k∂u k dt ;here the chain rule has been used to rewrite the last term on the right-h<strong>and</strong> side.Using (26.75) to write the derivatives of the basis vectors in terms of Christoffelsymbols, we obtaindvdt = dvidt e i +Γ j dukikvi dt e j.Interchanging the dummy indices i <strong>and</strong> j in the last term, we may factor out thebasis vector <strong>and</strong> find( )dv dvidt = dt +Γi jkv j duk e i .dtThe term in parentheses is called the absolute (or intrinsic) derivative of thecomponents v i along the curve r(t)<strong>and</strong> is usually denoted byδv iδt ≡ dvidt +Γi jkv j duk = v i du k; kdt dt .With this notation, we may writedvdt = δviδt e i = v i du k; kdt e i. (26.99)Using the same method, the absolute derivative of the covariant componentsv i of a vector is given byδv iδt ≡ v du ki; kdt .Similarly, the absolute derivatives of the contravariant, mixed <strong>and</strong> covariant975


TENSORScomponents of a second-order tensor T areδT ij≡ T ij du k; kδt dt ,δT i j≡ T i du kj; kδt dt ,δT ij du k≡ T ij; kδt dt .The derivative of T along the curve r(t) may then be written in terms of, <strong>for</strong>example, its contravariant components asdTdtijδT=δt e i ⊗ e j = T ij du k; kdt e i ⊗ e j .26.22 GeodesicsAs an example of the use of the absolute derivative, we conclude this chapterwith a brief discussion of geodesics. A geodesic in real three-dimensional spaceis a straight line, which has two equivalent defining properties. Firstly, it is thecurve of shortest length between two points <strong>and</strong>, secondly, it is the curve whosetangent vector always points in the same direction (along the line). Althoughin this chapter we have considered explicitly only our familiar three-dimensionalspace, much of the mathematical <strong>for</strong>malism developed can be generalised to moreabstract spaces of higher dimensionality in which the familiar ideas of Euclideangeometry are no longer valid. It is often of interest to find geodesic curves insuch spaces by using the defining properties of straight lines in Euclidean space.We shall not consider these more complicated spaces explicitly but will determinethe equation that a geodesic in Euclidean three-dimensional space (i.e.a straight line) must satisfy, deriving it in a sufficiently general way that ourmethod may be applied with little modification to finding the equations satisfiedby geodesics in more abstract spaces.Let us consider a curve r(s), parameterised by the arc length s from some pointon the curve, <strong>and</strong> choose as our defining property <strong>for</strong> a geodesic that its tangentvector t = dr/ds always points in the same direction everywhere on the curve, i.e.dt= 0. (26.100)dsAlternatively, we could exploit the property that the distance between two pointsis a minimum along a geodesic <strong>and</strong> use the calculus of variations (see chapter 22);this would lead to the same final result (26.101).If we now introduce an arbitrary coordinate system u i with basis vectors e i ,i =1, 2, 3, then we may write t = t i e i , <strong>and</strong> from (26.99) we finddtds = du kti ; kds e i = 0.976


26.23 EXERCISESWriting out the covariant derivative, we obtain( )dtids +Γi jkt j duk e i = 0.dsBut, since t j = du j /ds, it follows that the equation satisfied by a geodesic isd 2 u ids 2du j du k+Γi jk =0. (26.101)ds ds◮Find the equations satisfied by a geodesic (straight line) in cylindrical polar coordinates.From (26.83), the only non-zero Christoffel symbols are Γ 1 22 = −ρ <strong>and</strong> Γ2 12 =Γ2 21 =1/ρ.Thus the required geodesic equations are( )d 2 u 1 du 2 du 2ds 2 +Γ1 22ds ds =0 ⇒ d 2 2ρ dφds − ρ =0,2 dsd 2 u 2ds 2 +2Γ2 12du 1 du 2ds ds =0 ⇒ d 2 φds + 2 2 ρd 2 u 3ds 2 =0 ⇒d 2 zds =0. ◭2dρ dφds ds =0,26.23 Exercises26.1 Use the basic definition of a Cartesian tensor to show the following.(a) That <strong>for</strong> any general, but fixed, φ,(u 1 ,u 2 )=(x 1 cos φ − x 2 sin φ, x 1 sin φ + x 2 cos φ)are the components of a first-order tensor in two dimensions.(b) That( )x22 x 1 x 2x 1 x 2 x 2 1is not a tensor of order 2. To establish that a single element does nottrans<strong>for</strong>m correctly is sufficient.26.2 The components of two vectors, A <strong>and</strong> B, <strong>and</strong> a second-order tensor, T, are givenin one coordinate system by⎛A = ⎝ 1 ⎞ ⎛0 ⎠ , B = ⎝ 0 ⎞ ⎛1 ⎠ , T = ⎝ 2 √ ⎞√ 3 03 4 0 ⎠ .000 0 2In a second coordinate system, obtained from the first by rotation, the componentsof A <strong>and</strong> B are⎛A ′ = 1 ⎝2√301⎞ ⎛⎠ , B ′ = 1 ⎝2⎞−1√0 ⎠ .3Find the components of T in this new coordinate system <strong>and</strong> hence evaluate,with a minimum of calculation,T ij T ji , T ki T jk T ij , T ik T mn T ni T km .977


TENSORS26.3 In section 26.3 the trans<strong>for</strong>mation matrix <strong>for</strong> a rotation of the coordinate axeswas derived, <strong>and</strong> this approach is used in the rest of the chapter. An alternativeview is that of taking the coordinate axes as fixed <strong>and</strong> rotating the componentsof the system; this is equivalent to reversing the signs of all rotation angles.Using this alternative view, determine the matrices representing (a) a positiverotation of π/4 about the x-axis <strong>and</strong> (b) a rotation of −π/4 about the y-axis.Determine the initial vector r which, when subjected to (a) followed by (b),finishes at (3, 2, 1).26.4 Show how to decompose the Cartesian tensor T ij into three tensors,T ij = U ij + V ij + S ij ,where U ij is symmetric <strong>and</strong> has zero trace, V ij is isotropic <strong>and</strong> S ij has only threeindependent components.26.5 Use the quotient law discussed in section 26.7 to show that the array⎛⎞⎝ y2 + z 2 − x 2 −2xy −2xz−2yx x 2 + z 2 − y 2 −2yz ⎠−2zx −2zy x 2 + y 2 − z 2<strong>for</strong>ms a second-order tensor.26.6 Use tensor methods to establish the following vector identities:(a) (u × v) × w =(u · w)v − (v · w)u;(b) curl (φu) =φ curl u + (grad φ) × u;(c) div (u × v) =v · curl u − u · curl v;(d) curl (u × v) =(v · grad)u − (u · grad)v + u div v − v div u;(e) grad 1 (u · u) =u × curl u +(u · grad)u.226.7 Use result (e) of the previous question <strong>and</strong> the general divergence theorem <strong>for</strong>tensors to show that, <strong>for</strong> a vector field A,∫S[A(A · dS) −12 A2 dS ] =∫V[A divA − A × curl A] dV ,where S is the surface enclosing volume V .26.8 A column matrix a has components a x ,a y ,a z <strong>and</strong> A is the matrix with elementsA ij = −ɛ ijk a k .(a) What is the relationship between column matrices b <strong>and</strong> c if Ab = c?(b) Find the eigenvalues of A <strong>and</strong> show that a is one of its eigenvectors. Explainwhy this must be so.26.9 Equation (26.29),|A|ɛ lmn = A li A mj A nk ɛ ijk ,is a more general <strong>for</strong>m of the expression (8.47) <strong>for</strong> the determinant of a 3 × 3matrix A. The latter could have been written as|A| = ɛ ijk A i1 A j2 A k3 ,whilst the <strong>for</strong>mer removes the explicit mention of 1, 2, 3 at the expense of anadditional Levi–Civita symbol. As stated in the footnote on p. 942, (26.29) canbereadilyextendedtocoverageneralN × N matrix.Use the <strong>for</strong>m given in (26.29) to prove properties (i), (iii), (v), (vi) <strong>and</strong> (vii)of determinants stated in subsection 8.9.1. Property (iv) is obvious by inspection.For definiteness take N = 3, but convince yourself that your methods of proofwould be valid <strong>for</strong> any positive integer N.978


26.23 EXERCISES26.10 A symmetric second-order Cartesian tensor is defined byT ij = δ ij − 3x i x j .Evaluate the following surface integrals, each taken over the surface of the unitsphere:∫∫∫(a) T ij dS; (b) T ik T kj dS; (c) x i T jk dS.26.11 Given a non-zero vector v, find the value that should be assigned to α to makeP ij = αv i v j <strong>and</strong> Q ij = δ ij − αv i v jinto parallel <strong>and</strong> orthogonal projection tensors, respectively, i.e. tensors thatsatisfy, respectively, P ij v j = v i , P ij u j =0<strong>and</strong>Q ij v j =0,Q ij u j = u i ,<strong>for</strong>anyvectoru that is orthogonal to v.Show, in particular, that Q ij is unique, i.e. that if another tensor T ij has thesame properties as Q ij then (Q ij − T ij )w j =0<strong>for</strong>any vector w.26.12 In four dimensions, define second-order antisymmetric tensors, F ij <strong>and</strong> Q ij ,<strong>and</strong>a first-order tensor, S i , as follows:(a) F 23 = H 1 , Q 23 = B 1 <strong>and</strong> their cyclic permutations;(b) F i4 = −D i , Q i4 = E i <strong>for</strong> i =1, 2, 3;(c) S 4 = ρ, S i = J i <strong>for</strong> i =1, 2, 3.Then, taking x 4 as t <strong>and</strong> the other symbols to have their usual meanings inelectromagnetic theory, show that the equations ∑ j ∂F ij/∂x j = S i <strong>and</strong> ∂Q jk /∂x i +∂Q ki /∂x j + ∂Q ij /∂x k = 0 reproduce Maxwell’s equations. In the latter i, j, k is anyset of three subscripts selected from 1, 2, 3, 4, but chosen in such a way that theyare all different.26.13 In a certain crystal the unit cell can be taken as six identical atoms lying at thecorners of a regular octahedron. Convince yourself that these atoms can also beconsidered as lying at the centres of the faces of a cube <strong>and</strong> hence that the crystalhas cubic symmetry. Use this result to prove that the conductivity tensor <strong>for</strong> thecrystal, σ ij ,mustbeisotropic.26.14 Assuming that the current density j <strong>and</strong> the electric field E appearing in equation(26.44) are first-order Cartesian tensors, show explicitly that the electrical conductivitytensor σ ij trans<strong>for</strong>ms according to the law appropriate to a second-ordertensor.The rate W at which energy is dissipated per unit volume, as a result of thecurrent flow, is given by E · j. Determine the limits between which W must lie <strong>for</strong>a given value of |E| as the direction of E is varied.26.15 In a certain system of units, the electromagnetic stress tensor M ij is given byM ij = E i E j + B i B j − 1 δ 2 ij(E k E k + B k B k ),where the electric <strong>and</strong> magnetic fields, E <strong>and</strong> B, are first-order tensors. Show thatM ij is a second-order tensor.Consider a situation in which |E| = |B|, but the directions of E <strong>and</strong> B arenot parallel. Show that E ± B are principal axes of the stress tensor <strong>and</strong> findthe corresponding principal values. Determine the third principal axis <strong>and</strong> itscorresponding principal value.26.16 A rigid body consists of four particles of masses m, 2m, 3m, 4m, respectivelysituated at the points (a, a, a), (a, −a, −a), (−a, a, −a), (−a, −a, a) <strong>and</strong> connectedtogether by a light framework.(a) Find the inertia tensor at the origin <strong>and</strong> show that the principal moments ofinertia are 20ma 2 <strong>and</strong> (20 ± 2 √ 5)ma 2 .979


TENSORS(b) Find the principal axes <strong>and</strong> verify that they are orthogonal.26.17 A rigid body consists of eight particles, each of mass m, held together by lightrods. In a certain coordinate frame the particles are at positions±a(3, 1, −1), ±a(1, −1, 3), ±a(1, 3, −1), ±a(−1, 1, 3).Show that, when the body rotates about an axis through the origin, if the angularvelocity <strong>and</strong> angular momentum vectors are parallel then their ratio must be40ma 2 ,64ma 2 or 72ma 2 .26.18 The paramagnetic tensor χ ij of a body placed in a magnetic field, in which itsenergy density is − 1 µ 2 0M · H with M i = ∑ j χ ijH j ,is⎛⎝ 2k 0 0 3k 0k⎞⎠ .0 k 3kAssuming depolarizing effects are negligible, find how the body will orientateitself if the field is horizontal, in the following circumstances:(a) the body can rotate freely;(b) the body is suspended with the (1, 0, 0) axis vertical;(c) the body is suspended with the (0, 1, 0) axis vertical.26.19 A block of wood contains a number of thin soft-iron nails (of constant permeability).A unit magnetic field directed eastwards induces a magnetic moment inthe block having components (3, 1, −2), <strong>and</strong> similar fields directed northwards<strong>and</strong> vertically upwards induce moments (1, 3, −2) <strong>and</strong> (−2, −2, 2) respectively.Show that all the nails lie in parallel planes.26.20 For tin, the conductivity tensor is diagonal, with entries a, a, <strong>and</strong> b when referredto its crystal axes. A single crystal is grown in the shape of a long wire of length L<strong>and</strong> radius r, the axis of the wire making polar angle θ with respect to the crystal’s3-axis. Show that the resistance of the wire is L(πr 2 ab) ( −1 a cos 2 θ + b sin 2 θ ) .26.21 By considering an isotropic body subjected to a uni<strong>for</strong>m hydrostatic pressure(no shearing stress), show that the bulk modulus k, defined by the ratio of thepressure to the fractional decrease in volume, is given by k = E/[3(1−2σ)] whereE is Young’s modulus <strong>and</strong> σ is Poisson’s ratio.26.22 For an isotropic elastic medium under dynamic stress, at time t the displacementu i <strong>and</strong> the stress tensor p ij satisfy( ∂ukp ij = c ijkl + ∂u )l∂x l ∂x k<strong>and</strong>∂p ij= ρ ∂2 u i∂x j ∂t , 2where c ijkl is the isotropic tensor given in equation (26.47) <strong>and</strong> ρ is a constant.Show that both ∇ · u <strong>and</strong> ∇×u satisfy wave equations <strong>and</strong> find the correspondingwave speeds.980


26.23 EXERCISES26.23 A fourth-order tensor T ijkl has the propertiesT jikl = −T ijkl , T ijlk = −T ijkl .Prove that <strong>for</strong> any such tensor there exists a second-order tensor K mn such thatT ijkl = ɛ ijm ɛ kln K mn<strong>and</strong> give an explicit expression <strong>for</strong> K mn . Consider two (separate) special cases, asfollows.(a) Given that T ijkl is isotropic <strong>and</strong> T ijji = 1, show that T ijkl is uniquely determined<strong>and</strong> express it in terms of Kronecker deltas.(b) If now T ijkl has the additional propertyT klij = −T ijkl ,show that T ijkl has only three linearly independent components <strong>and</strong> find anexpression <strong>for</strong> T ijkl in terms of the vectorV i = − 1 ɛ 4 jklT ijkl .26.24 Working in cylindrical polar coordinates ρ, φ, z, parameterise the straight line(geodesic) joining (1, 0, 0) to (1,π/2, 1) in terms of s, the distance along the line.Show by substitution that the geodesic equations, derived at the end of section26.22, are satisfied.26.25 In a general coordinate system u i , i =1, 2, 3, in three-dimensional Euclideanspace, a volume element is given bydV = |e 1 du 1 · (e 2 du 2 × e 3 du 3 )|.Show that an alternative <strong>for</strong>m <strong>for</strong> this expression, written in terms of the determinantg of the metric tensor, is given bydV = √ gdu 1 du 2 du 3 .Show that, under a general coordinate trans<strong>for</strong>mation to a new coordinatesystem u ′i , the volume element dV remains unchanged, i.e. show that it is a scalarquantity.26.26 By writing down the expression <strong>for</strong> the square of the infinitesimal arc length (ds) 2in spherical polar coordinates, find the components g ij ofthemetrictensorinthiscoordinate system. Hence, using (26.97), find the expression <strong>for</strong> the divergenceof a vector field v in spherical polars. Calculate the Christoffel symbols (of thesecond kind) Γ i jkin this coordinate system.26.27 Find an expression <strong>for</strong> the second covariant derivative v i; jk ≡ (v i; j ) ; k of a vectorv i (see (26.88)). By interchanging the order of differentiation <strong>and</strong> then subtractingthe two expressions, we define the components R l ijkof the Riemann tensor asv i; jk − v i; kj ≡ R l ijkv l .Show that in a general coordinate system u i these components are given byR l ijk = ∂Γl ik∂u j− ∂Γl ij∂u k+Γ m ikΓ l mj − Γ m ijΓ l mk.By first considering Cartesian coordinates, show that all the components R l ijk ≡ 0<strong>for</strong> any coordinate system in three-dimensional Euclidean space.In such a space, there<strong>for</strong>e, we may change the order of the covariant derivativeswithout changing the resulting expression.981


TENSORS26.28 A curve r(t) is parameterised by a scalar variable t. Show that the length of thecurve between two points, A <strong>and</strong> B, isgivenby∫ √ BduL = g i du jijA dt dt dt.Using the calculus of variations (see chapter 22), show that the curve r(t) thatminimises L satisfies the equationd 2 u idt 2du j du k+Γi jk = ¨ṡ du idt dt s dt ,where s is the arc length along the curve, ṡ = ds/dt <strong>and</strong> ¨s = d 2 s/dt 2 . Hence, showthat if the parameter t is of the <strong>for</strong>m t = as + b, wherea <strong>and</strong> b are constants,then we recover the equation <strong>for</strong> a geodesic (26.101).[ A parameter which, like t, is the sum of a linear trans<strong>for</strong>mation of s <strong>and</strong> atranslation is called an affine parameter. ]26.29 We may define Christoffel symbols of the first kind byΓ ijk = g il Γ l jk.Show that these are given byΓ kij = 1 2By permuting indices, verify that(∂gik∂u j+ ∂g jk∂u i− ∂g ij∂u k ).∂g ij∂u =Γ k ijk +Γ jik .Using the fact that Γ l jk =Γl kj, show thatg ij; k ≡ 0,i.e. that the covariant derivative of the metric tensor is identically zero in allcoordinate systems.26.24 Hints <strong>and</strong> answers26.1 (a) u ′ 1 = x 1 cos(φ − θ) − x 2 sin(φ − θ), etc.;(b) u ′ 11 = s2 x 2 1 − 2scx 1x 2 + c 2 x 2 2 ≠ c2 x 2 2 + csx 1x 2 + scx 1 x 2 + s 2 x 2 1 .26.3 (a) (1/ √ 2)( √ 2, 0, 0; 0, 1, −1; 0, 1, 1). (b) (1/ √ 2)(1, 0, −1; 0, √ 2, 0; 1, 0, 1).r =(2 √ 2, −1+ √ 2, −1 − √ 2) T .26.5 Twice contract the array with the outer product of (x, y, z) with itself, thusobtaining the expression −(x 2 + y 2 + z 2 ) 2 , which is an invariant <strong>and</strong> there<strong>for</strong>e ascalar.26.7 Write A j (∂A i /∂x j )as∂(A i A j )/∂x j − A i (∂A j /∂x j ).26.9 (i) Write out the expression <strong>for</strong> |A T |, contract both sides of the equation with ɛ lmn<strong>and</strong> pick out the expression <strong>for</strong> |A| ontheRHS.Notethatɛ lmn ɛ lmn is a numericalscalar.(iii) Each non-zero term on the RHS contains any particular row index once <strong>and</strong>only once. The same can be said <strong>for</strong> the Levi–Civita symbol on the LHS. Thusinterchanging two rows is equivalent to interchanging two of the subscripts ofɛ lmn , <strong>and</strong> thereby reversing its sign. Consequently, the magnitude of |A| remainsthe same but its sign is changed.(v) If, say, A pi = λA pj , <strong>for</strong> some particular pair of values i <strong>and</strong> j <strong>and</strong> all p then,982


26.24 HINTS AND ANSWERSin the (multiple) summation on the RHS, each A nk appears multiplied by (withno summation over i <strong>and</strong> j)ɛ ijk A li A mj + ɛ jik A lj A mi = ɛ ijk λA lj A mj + ɛ jik A lj λA mj =0,since ɛ ijk = −ɛ jik . Consequently, grouped in this way all terms are zero <strong>and</strong>|A| =0.(vi) Replace A mj by A mj + λA lj <strong>and</strong> note that λA li A lj A nk ɛ ijk = 0 by virtue ofresult (v).(vii) If C = AB,|C|ɛ lmn = A lx B xi A my B yj A nz B zk ɛ ijk .Contract this with ɛ lmn <strong>and</strong> show that the RHS is equal to ɛ xyz |A T |ɛ xyz |B|. Itthenfollows from result (i) that |C| = |A||B|.26.11 α = |v| −2 . Note that the most general vector has components w i = λv i +µu (1)i+νu (2)i,where both u (1) <strong>and</strong> u (2) are orthogonal to v.26.13 Construct the orthogonal trans<strong>for</strong>mation matrix S <strong>for</strong> the symmetry operationof (say) a rotation of 2π/3 about a body diagonal <strong>and</strong>, setting L = S −1 = S T ,construct σ ′ = LσL T <strong>and</strong> require σ ′ = σ. Repeat the procedure <strong>for</strong> (say) a rotationof π/2 about the x 3 -axis. These together show that σ 11 = σ 22 = σ 33 <strong>and</strong> thatall other σ ij = 0. Further symmetry requirements do not provide any additionalconstraints.26.15 The trans<strong>for</strong>mation of δ ij has to be included; the principal values are ±E · B.The third axis is in the direction ±B × E with principal value −|E| 2 .26.17 The principal moments give the required ratios.26.19 The principal permeability, in direction (1, 1, 2), has value 0. Thus all the nails liein planes to which this is the normal.26.21 Take p 11 = p 22 = p 33 = −p, <strong>and</strong> p ij = e ij =0<strong>for</strong>i ≠ j, leading to −p =(λ+2µ/3)e ii . The fractional volume change is e ii ; λ <strong>and</strong> µ are as defined in (26.46)<strong>and</strong> the worked example that follows it.26.23 Consider Q pq = ɛ pij ɛ qkl T ijkl <strong>and</strong> show that K mn = Q mn /4 has the required property.(a) Argue from the isotropy of T ijkl <strong>and</strong> ɛ ijk <strong>for</strong> that of K mn <strong>and</strong> hence that itmust be a multiple of δ mn . Show that the multiplier is uniquely determined <strong>and</strong>that T ijkl =(δ il δ jk − δ ik δ jl )/6.(b) By relabelling dummy subscripts <strong>and</strong> using the stated antisymmetry property,show that K nm = −K mn . Show that −2V i = ɛ min K mn <strong>and</strong> hence that K mn = ɛ imn V i .26.25T ijkl = ɛ kli V j − ɛ klj V i .Use |e 1 · (e 2 × e 3 )| = √ g.Recall that √ g ′ = |∂u/∂u ′ | √ g <strong>and</strong> du ′1 du ′2 du ′3 = |∂u ′ /∂u| du 1 du 2 du 3 .26.27 (v i; j ) ; k =(v i; j ) ,k − Γ l ik v l; j − Γ l jk v i; l <strong>and</strong> v i; j = v i, j − Γ m ijv m . If all components of atensor equal zero in one coordinate system, then they are zero in all coordinatesystems.26.29 Use g il g ln = δi n <strong>and</strong> g ij = g ji . Show that( )∂gijg ij;k =∂u − Γ k jik − Γ ijk e i ⊗ e j<strong>and</strong>thenusetheearlierresult.983


27Numerical methodsIt happens frequently that the end product of a calculation or piece of analysisis one or more algebraic or differential equations, or an integral that cannot beevaluated in closed <strong>for</strong>m or in terms of tabulated or pre-programmed functions.From the point of view of the physical scientist or engineer, who needs numericalvalues <strong>for</strong> prediction or comparison with experiment, the calculation or analysisis thus incomplete.With the ready availability of st<strong>and</strong>ard packages on powerful computers <strong>for</strong>the numerical solution of equations, both algebraic <strong>and</strong> differential, <strong>and</strong> <strong>for</strong> theevaluation of integrals, in principle there is no need <strong>for</strong> the investigator to doanything other than turn to them. However, it should be a part of every engineer’sor scientist’s repertoire to have some underst<strong>and</strong>ing of the kinds of procedure thatare being put into practice within those packages. The present chapter indicates(at a simple level) some of the ways in which analytically intractable problemscan be tackled using numerical methods.In the restricted space available in a book of this nature, it is clearly notpossible to give anything like a full discussion, even of the elementary points thatwill be made in this chapter. The limited objective adopted is that of explaining<strong>and</strong> illustrating by simple examples some of the basic principles involved. Inmany cases, the examples used can be solved in closed <strong>for</strong>m anyway, but this‘obviousness’ of the answers should not detract from their illustrative usefulness,<strong>and</strong> it is hoped that their transparency will help the reader to appreciate some ofthe inner workings of the methods described.The student who proposes to study complicated sets of equations or makerepeated use of the same procedures by, <strong>for</strong> example, writing computer programsto carry out the computations, will find it essential to acquire a good underst<strong>and</strong>ingof topics hardly mentioned here. Amongst these are the sensitivity ofthe adopted procedures to errors introduced by the limited accuracy with whicha numerical value can be stored in a computer (rounding errors) <strong>and</strong> to the984


27.1 ALGEBRAIC AND TRANSCENDENTAL EQUATIONSerrors introduced as a result of approximations made in setting up the numericalprocedures (truncation errors). For this scale of application, books specificallydevoted to numerical analysis, data analysis <strong>and</strong> computer programming shouldbe consulted.So far as is possible, the method of presentation here is that of indicating<strong>and</strong> discussing in a qualitative way the main steps in the procedure, <strong>and</strong> thenof following this with an elementary worked example. The examples have beenrestricted in complexity to a level at which they can be carried out with a pocketcalculator. Naturally it will not be possible <strong>for</strong> the student to check all thenumerical values presented, unless he or she has a programmable calculator orcomputer readily available, <strong>and</strong> even then it might be tedious to do so. However,it is advisable to check the initial step <strong>and</strong> at least one step in the middle ofeach repetitive calculation given in the text, so that how the symbolic equationsare used with actual numbers is understood. Clearly the intermediate step shouldbe chosen to be at a point in the calculation at which the changes are stillsufficiently large that they can be detected by whatever calculating device isused.Where alternative methods <strong>for</strong> solving the same type of problem are discussed,<strong>for</strong> example in finding the roots of a polynomial equation, we have usuallytaken the same example to illustrate each method. This could give the mistakenimpression that the methods are very restricted in applicability, but it is felt bythe authors that using the same examples repeatedly has sufficient advantages, interms of illustrating the relative characteristics of competing methods, to justifydoing so. Once the principles are clear, little is to be gained by using newexamples each time, <strong>and</strong>, in fact, having some prior knowledge of the ‘correctanswer’ should allow the reader to judge the efficiency <strong>and</strong> dangers of particularmethods as the successive steps are followed through.One other point remains to be mentioned. Here, in contrast with every otherchapter of this book, the value of a large selection of exercises is not clear cut.The reader with sufficient computing resources to tackle them can easily devisealgebraic or differential equations to be solved, or functions to be integrated(which perhaps have arisen in other contexts). Further, the solutions of theseproblems will be self-checking, <strong>for</strong> the most part. Consequently, although anumber of exercises are included, no attempt has been made to test the full rangeof ideas treated in this chapter.27.1 Algebraic <strong>and</strong> transcendental equationsThe problem of finding the real roots of an equation of the <strong>for</strong>m f(x) =0,wheref(x) is an algebraic or transcendental function of x, is one that can sometimesbe treated numerically, even if explicit solutions in closed <strong>for</strong>m are not feasible.985


NUMERICAL METHODSf(x)141210 f(x) =x 5 − 2x 2 − 3864200.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6x1.8−2−4Figure 27.10 ≤ x ≤ 1.9.A graph of the function f(x) =x 5 − 2x 2 − 3<strong>for</strong>x in the rangeExamples of the types of equation mentioned are the quartic equation,<strong>and</strong> the transcendental equation,ax 4 + bx + c =0,x − 3 tanh x =0.The latter type is characterised by the fact that it contains, in effect, a polynomialof infinite order on the LHS.We will discuss four methods that, in various circumstances, can be used toobtain the real roots of equations of the above types. In all cases we will take asthe specific equation to be solved the fifth-order polynomial equationf(x) ≡ x 5 − 2x 2 − 3=0. (27.1)The reasons <strong>for</strong> using the same equation each time were discussed in the introductionto this chapter.For future reference, <strong>and</strong> so that the reader may follow some of the calculationsleading to the evaluation of the real root of (27.1), a graph of f(x) in the range0 ≤ x ≤ 1.9 is shown in figure 27.1.Equation (27.1) is one <strong>for</strong> which no solution can be found in closed <strong>for</strong>m, thatis in the <strong>for</strong>m x = a, wherea does not explicitly contain x. The general scheme tobe employed will be an iterative one in which successive approximations to a realroot of (27.1) will be obtained, each approximation, it is to be hoped, being betterthan the preceding one; certainly, we require that the approximations converge<strong>and</strong> that they have as their limit the sought-<strong>for</strong> root. Let us denote the required986


27.1 ALGEBRAIC AND TRANSCENDENTAL EQUATIONSroot by ξ <strong>and</strong> the values of successive approximations by x 1 , x 2 , ..., x n , ... . Then,<strong>for</strong> any particular method to be successful,lim x n = ξ, where f(ξ) =0. (27.2)n→∞However, success as defined here is not the only criterion. Since, in practice,only a finite number of iterations will be possible, it is important that the valuesof x n be close to that of ξ <strong>for</strong> all n>N,whereN is a relatively low number;exactly how low it is naturally depends on the computing resources available <strong>and</strong>the accuracy required in the final answer.So that the reader may assess the progress of the calculations that follow, werecord that to nine significant figures the real root of equation (27.1) has thevalueξ =1.495 106 40. (27.3)We now consider in turn four methods <strong>for</strong> determining the value of this root.27.1.1 Rearrangement of the equationIf equation (27.1), f(x) = 0, can be recast into the <strong>for</strong>mx = φ(x), (27.4)where φ(x) isaslowly varying function of x, then an iteration schemex n+1 = φ(x n ) (27.5)will often produce a fair approximation to the root ξ after a few iterations,as follows. Clearly, ξ = φ(ξ), since f(ξ) = 0; thus, when x n is close to ξ,the next approximation, x n+1 , will differ little from x n , the actual size of thedifference giving an order-of-magnitude indication of the inaccuracy in x n+1(when compared with ξ).In the present case, the equation can be writtenx =(2x 2 +3) 1/5 . (27.6)Because of the presence of the one-fifth power, the RHS is rather insensitiveto the value of x used to compute it, <strong>and</strong> so the <strong>for</strong>m (27.6) fits the generalrequirements <strong>for</strong> the method to work satisfactorily. It remains only to choose astarting approximation. It is easy to see from figure 27.1 that the value x =1.5would be a good starting point, but, so that the behaviour of the procedure atvalues some way from the actual root can be studied, we will make a poorerchoice, x 1 =1.7.With this starting value <strong>and</strong> the general recurrence relationshipx n+1 =(2x 2 n +3) 1/5 , (27.7)987


NUMERICAL METHODSn x n f(x n )1 1.7 5.422 1.544 18 1.013 1.506 86 2.28 × 10 −14 1.497 92 5.37 × 10 −25 1.495 78 1.28 × 10 −26 1.495 27 3.11 × 10 −37 1.495 14 7.34 × 10 −48 1.495 12 1.76 × 10 −4Table 27.1 Successive approximations to the root of (27.1) using the methodof rearrangement.n A n f(A n ) B n f(B n ) x n f(x n )1 1.0 −4.0000 1.7 5.4186 1.2973 −2.69162 1.2973 −2.6916 1.7 5.4186 1.4310 −1.09573 1.4310 −1.0957 1.7 5.4186 1.4762 −0.34824 1.4762 −0.3482 1.7 5.4186 1.4897 −0.10165 1.4897 −0.1016 1.7 5.4186 1.4936 −0.02896 1.4936 −0.0289 1.7 5.4186 1.4947 −0.0082Table 27.2 Successive approximations to the root of (27.1) using linearinterpolation.successive values can be found. These are recorded in table 27.1. Although notstrictly necessary, the value of f(x n ) ≡ x 5 n − 2x 2 n − 3 is also shown at each stage.It will be seen that x 7 <strong>and</strong> all later x n agree with the precise answer (27.3) towithin one part in 10 4 . However, f(x n )<strong>and</strong>x n − ξ are both reduced by a factorof only about 4 <strong>for</strong> each iteration; thus a large number of iterations would beneeded to produce a very accurate answer. The factor 4 is, of course, specificto this particular problem <strong>and</strong> would be different <strong>for</strong> a different equation. Thesuccessive values of x n are shown in graph (a) of figure 27.2.27.1.2 Linear interpolationIn this approach two values, A 1 <strong>and</strong> B 1 ,ofx are chosen with A 1


27.1 ALGEBRAIC AND TRANSCENDENTAL EQUATIONS6(a)6(b)44x 1x 22x 22−2−4642−2−41.0 1.2 1.4 1.6ξ(c)x 3x 3x 14x 3 x 42ξ1.0 1.2 1.4x1.61.0 1.2 1.4 1.62x2−2ξx 1−41.0−2−4(d)61.2x 11.4x 3ξ1.6Figure 27.2 Graphical illustrations of the iteration methods discussed inthe text: (a) rearrangement; (b) linear interpolation; (c) binary chopping;(d) Newton–Raphson.with n =1.Next,f(x 1 ) is evaluated <strong>and</strong> the process repeated after replacingeither A 1 or B 1 by x 1 , according to whether f(x 1 ) has the same sign as f(A 1 )orf(B 1 ), respectively. In figure 27.2(b), A 1 is the one replaced.As can be seen in the particular example that we are considering, with thismethod there is a tendency, if the curvature of f(x) is of constant sign nearthe root, <strong>for</strong> one of the two ends of the successive chords to remain unchanged.Starting with the initial values A 1 = 1 <strong>and</strong> B 1 =1.7, the results of the firstfive iterations using (27.8) are given in table 27.2 <strong>and</strong> indicated in graph (b) offigure 27.2. As with the rearrangement method, the improvement in accuracy,as measured by f(x n )<strong>and</strong>x n − ξ, is a fairly constant factor at each iteration(approximately 3 in this case), <strong>and</strong> <strong>for</strong> our particular example there is little tochoose between the two. Both tend to their limiting value of ξ monotonically,from either higher or lower values, <strong>and</strong> this makes it difficult to estimate limitswithin which ξ can safely be presumed to lie. The next method to be describedgives at any stage a range of values within which ξ is known to lie.989


NUMERICAL METHODSn A n f(A n ) B n f(B n ) x n f(x n )1 1.0000 −4.0000 1.7000 5.4186 1.3500 −2.16102 1.3500 −2.1610 1.7000 5.4186 1.5250 0.59683 1.3500 −2.1610 1.5250 0.5968 1.4375 −0.99464 1.4375 −0.9946 1.5250 0.5968 1.4813 −0.25735 1.4813 −0.2573 1.5250 0.5968 1.5031 0.15446 1.4813 −0.2573 1.5031 0.1544 1.4922 −0.05527 1.4922 −0.0552 1.5031 0.1544 1.4977 0.04878 1.4922 −0.0552 1.4977 0.0487 1.4949 −0.0085Table 27.3chopping.Successive approximations to the root of (27.1) using binary27.1.3 Binary choppingAgain two values of x, A 1 <strong>and</strong> B 1 , that straddle the root are chosen, such thatA 1


27.1 ALGEBRAIC AND TRANSCENDENTAL EQUATIONSn x n f(x n )1 1.7 5.422 1.545 01 1.033 1.498 87 7.20 × 10 −24 1.495 13 4.49 × 10 −45 1.495 106 40 2.6 × 10 −86 1.495 106 40 —Table 27.4 Successive approximations to the root of (27.1) using the Newton–Raphson method.of (notionally) constructing the chord between two points on the curve of f(x)against x, the tangent to the curve is notionally constructed at each successivevalue of x n , <strong>and</strong> the next value, x n+1 , is taken as the point at which the tangentcuts the axis f(x) = 0. This is illustrated in graph (d) of figure 27.2.If the nth value is x n , the tangent to the curve of f(x) at that point has slopef ′ (x n ) <strong>and</strong> passes through the point x = x n , y = f(x n ). Its equation is thusy(x) =(x − x n )f ′ (x n )+f(x n ). (27.10)The value of x at which y = 0 is then taken as x n+1 ; thus the condition y(x n+1 )=0yields, from (27.10), the iteration schemex n+1 = x n − f(x n)f ′ (x n ) . (27.11)This is the Newton–Raphson iteration <strong>for</strong>mula. Clearly, if x n is close to ξ then x n+1is close to x n , as it should be. It is also apparent that if any of the x n comes closeto a stationary point of f, sothatf ′ (x n ) is close to zero, the scheme is not goingto work well.For our st<strong>and</strong>ard example, (27.11) becomesx n+1 = x n − x5 n − 2x 2 n − 35x 4 = 4x5 n − 2x 2 n +3n − 4x n 5x 4 . (27.12)n − 4x nAgain taking a starting value of x 1 =1.7, we obtain in succession the entriesin table 27.4. The different values are given to an increasing number of decimalplaces as the calculation proceeds; f(x n ) is also recorded.It is apparent that this method is unlike the previous ones in that the increasein accuracy of the answer is not constant throughout the iterations but improvesdramatically as the required root is approached. Away from the root the behaviourof the series is less satisfactory, <strong>and</strong> from its geometrical interpretation it can beseen that if, <strong>for</strong> example, there were a maximum or minimum near the root thenthe series could oscillate between values on either side of it (instead of ‘homingin’ on the root). The reason <strong>for</strong> the good convergence near the root is discussedin the next section.991


NUMERICAL METHODSOf the four methods mentioned, no single one is ideal, <strong>and</strong>, in practice, somemixture of them is usually to be preferred. The particular combination of methodsselected will depend a great deal on how easily the progress of the calculationmay be monitored, but some combination of the first three methods mentioned,followed by the NR scheme if great accuracy were required, would be suitable<strong>for</strong> most situations.27.2 Convergence of iteration schemesFor iteration schemes in which x n+1 can be expressed as a differentiable functionof x n , <strong>for</strong> example the rearrangement or NR methods of the previous section, apartial analysis of the conditions necessary <strong>for</strong> a successful scheme can be madeas follows.Suppose the general iteration <strong>for</strong>mula is expressed asx n+1 = F(x n ) (27.13)((27.7) <strong>and</strong> (27.12) are examples). Then the sequence of values x 1 ,x 2 ,...,x n ,... isrequired to converge to the value ξ that satisfies bothf(ξ) =0 <strong>and</strong> ξ = F(ξ). (27.14)If the error in the solution at the nth stage is ɛ n ,i.e.x n = ξ + ɛ n ,thenξ + ɛ n+1 = x n+1 = F(x n )=F(ξ + ɛ n ). (27.15)For the iteration process to converge, a decreasing error is required, i.e. |ɛ n+1 |


27.2 CONVERGENCE OF ITERATION SCHEMESy = xyx n x n+1 x n+2y = F(x)ξxFigure 27.3 Illustration of the convergence of the iteration scheme x n+1 =F(x n )when0


NUMERICAL METHODSn x n+1 ɛ n1 8.5 4.52 5.191 1.193 4.137 1.4 × 10 −14 4.002 257 2.3 × 10 −35 4.000 000 637 6.4 × 10 −76 4 —Table 27.5(27.20).Successive approximations to √ 16 using the iteration scheme◮The following is an iteration scheme <strong>for</strong> finding the square root of X:x n+1 = 1 (x n + X ). (27.20)2 x nShow that it has second-order convergence <strong>and</strong> illustrate its efficiency by finding, say, √ 16starting with a very poor guess, √ 16 = 1.If this scheme does converge to ξ then ξ will satisfyξ = 1 (ξ + X )⇒ ξ 2 = X,2 ξas required. The iteration function F is given byF(x) = 1 2(x + X x),<strong>and</strong> so, since ξ 2 = X,F ′ (ξ) = 1 2(1 − X x 2 )x=ξ=0,whilst( ) XF ′′ (ξ) = = 1 x 3 x=ξξ ≠0.Thus the procedure has second-order, but not third-order, convergence.We now show the procedure in action. Table 27.5 gives successive values of x n <strong>and</strong> ofɛ n , the difference between x n <strong>and</strong> the true value, 4. As we can see, the scheme is crudeinitially, but once x n gets close to ξ, it homes in on the true value extremely rapidly. ◭27.3 Simultaneous linear equationsAs we saw in chapter 8, many situations in physical science can be describedapproximately or exactly by a set of N simultaneous linear equations in N994


27.3 SIMULTANEOUS LINEAR EQUATIONSvariables (unknowns), x i , i =1, 2,...,N. The equations take the general <strong>for</strong>mA 11 x 1 + A 12 x 2 + ···+ A 1N x N = b 1 ,A 21 x 1 + A 22 x 2 + ···+ A 2N x N = b 2 , (27.21)..A N1 x 1 + A N2 x 2 + ···+ A NN x N = b N ,where the A ij are constants <strong>and</strong> <strong>for</strong>m the elements of a square matrix A. Theb iare given <strong>and</strong> <strong>for</strong>m a column matrix b. IfA is non-singular then (27.21) can besolved <strong>for</strong> the x i using the inverse of A, according to the <strong>for</strong>mulax = A −1 b.This approach was discussed at length in chapter 8 <strong>and</strong> will not be consideredfurther here.27.3.1 Gaussian eliminationWe follow instead a continuation of one of the earliest techniques acquired by astudent of algebra, namely the solving of simultaneous equations (initially onlytwo in number) by the successive elimination of all the variables but one. This(known as Gaussian elimination) is achieved by using, at each stage, one of theequations to obtain an explicit expression <strong>for</strong> one of the remaining x i in termsof the others <strong>and</strong> then substituting <strong>for</strong> that x i in all other remaining equations.Eventually a single linear equation in just one of the unknowns is obtained. Thisis then solved <strong>and</strong> the result is resubstituted in previously derived equations (inreverse order) to establish values <strong>for</strong> all the x i .This method is probably very familiar to the reader, <strong>and</strong> so a specific exampleto illustrate this alone seems unnecessary. Instead, we will show how a calculationalong such lines might be arranged so that the errors due to the inherent lack ofprecision in any calculating equipment do not become excessive. This can happenif the value of N is large <strong>and</strong> particularly (<strong>and</strong> we will merely state this) if theelements A 11 ,A 22 ,...,A NN on the leading diagonal of the matrix in (27.21) aresmall compared with the off-diagonal elements.The process to be described is known as Gaussian elimination with interchange.The only, but essential, difference from straight<strong>for</strong>ward elimination is that be<strong>for</strong>eeach variable x i is eliminated, the equations are reordered to put the largest (inmodulus) remaining coefficient of x i on the leading diagonal.We will take as an illustration a straight<strong>for</strong>ward three-variable example, whichcan in fact be solved perfectly well without any interchange since, with simplenumbers <strong>and</strong> only two eliminations to per<strong>for</strong>m, rounding errors do not havea chance to build up. However, the important thing is that the reader should995


NUMERICAL METHODSappreciate how this would apply in (say) a computer program <strong>for</strong> a 1000-variablecase, perhaps with un<strong>for</strong>seeable zeros or very small numbers appearing on theleading diagonal.◮Solve the simultaneous equations(a) x 1 +6x 2 −4x 3 = 8,(b) 3x 1 −20x 2 +x 3 = 12,(c) −x 1 +3x 2 +5x 3 = 3.(27.22)Firstly, we interchange rows (a) <strong>and</strong> (b) to bring the term 3x 1 onto the leading diagonal. Inthe following, we label the important equations (I), (II), (III), <strong>and</strong> the others alphabetically.A general (i.e. variable) label will be denoted by j.(I) 3x 1 −20x 2 +x 3 = 12,(d) x 1 +6x 2 −4x 3 = 8,(e) −x 1 +3x 2 +5x 3 = 3.For (j) = (d) <strong>and</strong> (e), replace row (j) byrow (j) − a j1× row (I),3where a j1 is the coefficient of x 1 in row (j), to give the two equations(II)( )6+20 x23+ ( )−4 − 1 x33= 8− 12 3(f)( )3 −20 x23+ ( )5+ 1 x33= 3+ 12 3Now |6+ 2020| > |3 − | <strong>and</strong> so no interchange is required be<strong>for</strong>e the next elimination. To3 3eliminate x 2 ,replacerow(f)by( )−113row (f) − × row (II).This gives[(III) 16] (−13) x3 =7+ 11 3 38 338Collecting together <strong>and</strong> tidying up the final equations, we have(I) 3x 1 −20x 2 +x 3 = 12,(II) 38x 2 −13x 3 = 12,(III) x 3 = 2.Starting with (III) <strong>and</strong> working backwards, it is now a simple matter to obtainx 1 =10, x 2 =1, x 3 =2. ◭38327.3.2 Gauss–Seidel iterationIn the example considered in the previous subsection an explicit way of solvinga set of simultaneous equations was given, the accuracy obtainable being limitedonly by the rounding errors in the calculating facilities available, <strong>and</strong> the calculationwas planned to minimise these. However, in some situations it may be thatonly an approximate solution is needed. If, <strong>for</strong> a large number of variables, this is996


27.3 SIMULTANEOUS LINEAR EQUATIONSthe case then an iterative method may produce a satisfactory degree of precisionwith less calculation. Such a method, known as Gauss–Seidel iteration, isbasedupon the following analysis.The problem is again that of finding the components of the column matrix xthat satisfiesAx = b (27.23)when A <strong>and</strong> b are a given matrix <strong>and</strong> column matrix, respectively.The steps of the Gauss–Seidel scheme are as follows.(i) Rearrange the equations (usually by simple division on both sides of eachequation) so that all diagonal elements of the new matrix C are unity, i.e.(27.23) becomesCx = d, (27.24)where C = I − F, <strong>and</strong>F has zeros as its diagonal elements.(ii) Step (i) produces<strong>and</strong> this <strong>for</strong>ms the basis of an iteration scheme,Fx + d = Ix = x, (27.25)x n+1 = Fx n + d, (27.26)where x n is the nth approximation to the required solution vector ξ.(iii) To improve the convergence, the matrix F, which has zeros on its leadingdiagonal, can be written as the sum of two matrices L <strong>and</strong> U that havenon-zero elements only below <strong>and</strong> above the leading diagonal, respectively:L ij =U ij ={Fij if i>j,0 otherwise,{Fij if i


NUMERICAL METHODSn x 1 x 2 x 31 2 2 22 4 0.1 1.343 12.76 1.381 2.3234 9.008 0.867 1.8815 10.321 1.042 2.0396 9.902 0.987 1.9887 10.029 1.004 2.004Table 27.6 Successive approximations to the solution of simultaneous equations(27.29) using the Gauss–Seidel iteration method.◮Obtain an approximate solution to the simultaneous equationsx 1 +6x 2 −4x 3 = 8,3x 1 −20x 2 +x 3 = 12,−x 1 +3x 2 +5x 3 = 3.These are the same equations as were solved in subsection 27.3.1.(27.29)Divide the equations by 1, −20 <strong>and</strong> 5, respectively, to givex 1 +6x 2 − 4x 3 =8,−0.15x 1 + x 2 − 0.05x 3 = −0.6,−0.2x 1 +0.6x 2 + x 3 =0.6.Thus, set out in matrix <strong>for</strong>m, (27.28) is, in this case, given by⎛⎝ x ⎞ ⎛1x 2⎠ = ⎝ 0 0 0⎞ ⎛0.15 0 0 ⎠ ⎝ x ⎞1x 2⎠x 30.2 −0.6 0 x 3n+1n+1⎛+ ⎝ 0 −6 4⎞ ⎛0 0 0.05 ⎠ ⎝ x ⎞ ⎛1x 2⎠ + ⎝ 8⎞−0.6 ⎠ .0 0 0 x 3 0.6nSuppose initially (n = 1) we guess each component to have the value 2. Then the successivesets of values of the three quantities generated by this scheme are as shown in table 27.6.Even with the rather poor initial guess, a close approximation to the exact result, x 1 = 10,x 2 =1,x 3 = 2, is obtained in only a few iterations. ◭27.3.3 Tridiagonal matricesAlthough <strong>for</strong> the solution of most matrix equations Ax = b the number ofoperations required increases rapidly with the size N × N of the matrix (roughlyas N 3 ), <strong>for</strong> one particularly simple kind of matrix the computing required increasesonly linearly with N. This type often occurs in physical situations in which objectsin an ordered set interact only with their nearest neighbours <strong>and</strong> is one in whichonly the leading diagonal <strong>and</strong> the diagonals immediately above <strong>and</strong> below it998


. . .. . .27.3 SIMULTANEOUS LINEAR EQUATIONScontain non-zero entries. Such matrices are known as tridiagonal matrices. Theymay also be used in numerical approximations to the solutions of certain typesof differential equation.A typical matrix equation involving a tridiagonal matrix is as follows:b 1c 10a 2c 3b 2c 2a 3b 3a N–1b N–1c N–10a Nb N. . .. . .. . .x 1y 1x 2y 2x 3y 3=x Nx N–1y N–1y N(27.30)So as to keep the entries in the matrix as free from subscripts as possible, wehave used a, b <strong>and</strong> c to indicate subdiagonal, leading diagonal <strong>and</strong> superdiagonalelements, respectively. As a consequence, we have had to change the notation <strong>for</strong>the column matrix on the RHS from b to (say) y.In such an equation the first <strong>and</strong> last rows involve x 1 <strong>and</strong> x N , respectively, <strong>and</strong>so the solution could be found by letting x 1 be unknown <strong>and</strong> then solving in turneach row of the equation in terms of x 1 , <strong>and</strong> finally determining x 1 by requiringthe next-to-last line to generate <strong>for</strong> x N an equation compatible with that givenby the last line. However, if the matrix is large this becomes a very cumbersomeoperation, <strong>and</strong> a simpler method is to assume a <strong>for</strong>m of solutionSince the ith line of the matrix equation isx i−1 = θ i−1 x i + φ i−1 . (27.31)a i x i−1 + b i x i + c i x i+1 = y i ,we must have, by substituting <strong>for</strong> x i−1 ,that(a i θ i−1 + b i )x i + c i x i+1 = y i − a i φ i−1 .This is also in the <strong>for</strong>m of (27.31), but with i replaced by i+1. Thus the recurrence<strong>for</strong>mulae <strong>for</strong> θ i <strong>and</strong> φ i areθ i =−c ia i θ i−1 + b i,φ i = y i − a i φ i−1a i θ i−1 + b i, (27.32)provided the denominator does not vanish <strong>for</strong> any i. From the first of the matrixequations it follows that θ 1 = −c 1 /b 1 <strong>and</strong> φ 1 = y 1 /b 1 . The equations may nowbe solved <strong>for</strong> the x i in two stages without carrying through an unknown quantity.First, all the θ i <strong>and</strong> φ i are generated using (27.32) <strong>and</strong> the values of θ 1 <strong>and</strong> φ 1 ,then, as a second stage, (27.31) is used to evaluate the x i , starting with x N (= φ N )<strong>and</strong> working backwards.999


NUMERICAL METHODS◮Solve the following tridiagonal matrix equation, in which only non-zero elements areshown:⎛⎞ ⎛ ⎞ ⎛ ⎞1 2x 1 4−1 2 1x 232 −1 2x ⎜ 3 1 13−3⎟ ⎜ x =4 ⎟ ⎜ 10. (27.33)⎟⎝3 4 2 ⎠ ⎝ x 5⎠ ⎝ 7 ⎠−2 2 x 6 −2The solution is set out in table 27.7, in which the arrows indicate the general flow of thecalculation. First, the columns of a i , b i , c i <strong>and</strong> y i are filled in from the original equation(27.33) <strong>and</strong> then the recurrence relations (27.32) are used to fill in the successive rowsstarting at the top; on each row we work from left to right as far as <strong>and</strong> including the φ icolumn. Finally, the bottom entry in the the x i column is set equal to the bottom entry inthe completed φ i column <strong>and</strong> the rest of the x i column is completed by using (27.31) <strong>and</strong>working up from the bottom. Thus the solution is x 1 =2;x 2 =1;x 3 =3;x 4 = −1; x 5 =a i b i c i a i θ i−1 + b i θ i y i a i φ i−1 φ i x i↓ 0 1 2 → 1 −2 4 0 4 2 ↑↓ −1 2 1 → 4 −1/4 3 −4 7/4 1 ↑↓ 2 −1 2 → −3/2 4/3 −3 7/2 13/3 3 ↑↓ 3 1 1 → 5 −1/5 10 13 −3/5 −1 ↑↓ 3 4 2 → 17/5 −10/17 7 −9/5 44/17 2 ↑↓ −2 2 0 → 54/17 0 −2 −88/17 1 → 1 ↑Table 27.7 The solution of tridiagonal matrix equation (27.33). The arrowsindicate the general flow of the calculation, as described in the text.2; x 6 =1.◭27.4 Numerical integrationAs noted at the start of this chapter, with modern computers <strong>and</strong> computerpackages – some of which will present solutions in algebraic <strong>for</strong>m, where thatis possible – the inability to find a closed-<strong>for</strong>m expression <strong>for</strong> an integral nolonger presents a problem. But, just as <strong>for</strong> the solution of algebraic equations, itis extremely important that scientists <strong>and</strong> engineers should have some idea of theprocedures on which such packages are based. In this section we discuss some ofthe more elementary methods used to evaluate integrals numerically <strong>and</strong> at thesame time indicate the basis of more sophisticated procedures.The st<strong>and</strong>ard integral evaluation has the <strong>for</strong>mI =∫ baf(x) dx, (27.34)where the integr<strong>and</strong> f(x) may be given in analytic or tabulated <strong>for</strong>m, but <strong>for</strong> thecases under consideration no closed-<strong>for</strong>m expression <strong>for</strong> I can be obtained. All1000


27.4 NUMERICAL INTEGRATION(a) (b) (c)f(x)f(x)f i+1f if i+1f i+1hhhhf if i−1x i+1x i x i+1/2 x i+1 x i+1x iFigure 27.4 (a) Definition of nomenclature. (b) The approximation in usingthe trapezium rule; f(x) is indicated by the broken curve. (c) Simpson’s ruleapproximation; f(x) is indicated by the broken curve. The solid curve is partof the approximating parabola.x i−1x inumerical evaluations of I are based on regarding I as the area under the curveof f(x) between the limits x = a <strong>and</strong> x = b <strong>and</strong> attempting to estimate that area.The simplest methods of doing this involve dividing up the interval a ≤ x ≤ binto N equal sections, each of length h =(b − a)/N. The dividing points arelabelled x i , with x 0 = a, x N = b, i running from 0 to N. The point x i is a distanceih from a. The central value of x in a strip (x = x i + h/2) is denoted <strong>for</strong> brevityby x i+1/2 , <strong>and</strong> <strong>for</strong> the same reason f(x i ) is written as f i . This nomenclature isindicated graphically in figure 27.4(a).So that we may compare later estimates of the area under the curve with thetrue value, we next obtain an exact expression <strong>for</strong> I, even though we cannotevaluate it. To do this we need to consider only one strip, say that between x i<strong>and</strong> x i+1 . For this strip the area is, using Taylor’s expansion,∫ h/2−h/2∫ h/2f(x i+1/2 + y) dy =−h/2∞∑==n=0∞∑∞∑f (n) (x i+1/2 ) ynn! dyn=0∫ h/2f (n)i+1/2f (n)i+1/2n even−h/2y nn! dy) n+12 h. (27.35)(n +1)!(2It should be noted that, in this exact expression, only the even derivatives off survive the integration <strong>and</strong> all derivatives are evaluated at x i+1/2 . Clearly1001


NUMERICAL METHODSother exact expressions are possible, e.g. the integral of f(x i + y) over the range0 ≤ y ≤ h, but we will find (27.35) the most useful <strong>for</strong> our purposes.Although the preceding discussion has implicitly assumed that both of thelimits a <strong>and</strong> b are finite, with the consequence that N is finite, the general methodcan be adapted to treat some cases in which one of the limits is infinite. It issufficient to consider one infinite limit, as an integral with limits −∞ <strong>and</strong> ∞ canbe considered as the sum of two integrals, each with one infinite limit.Consider the integralI =∫ ∞af(x) dx,where a is chosen large enough that the integr<strong>and</strong> is monotonically decreasing<strong>for</strong> x>a<strong>and</strong> falls off more quickly than x −2 . The change of variable t =1/xconverts this integral into∫ 1/a( )1 1I =0 t 2 f dt.tIt is now an integral over a finite range <strong>and</strong> the methods indicated earlier can beapplied to it. The value of the integr<strong>and</strong> at the lower end of the t-range is zero.In a similar vein, integrals with an upper limit of ∞ <strong>and</strong> an integr<strong>and</strong> that isknown to behave asymptotically as g(x)e −αx ,whereg(x) is a smooth function, canbe converted into an integral over a finite range by setting x = −α −1 ln αt. Again,the lower limit, a, <strong>for</strong> this part of the integral should be positive <strong>and</strong> chosenbeyond the last turning point of g(x). The part of the integral <strong>for</strong> x


27.4 NUMERICAL INTEGRATIONThis provides a very simple expression <strong>for</strong> estimating integral (27.34); its accuracyis limited only by the extent to which h can be made very small (<strong>and</strong> hence Nvery large) without making the calculation excessively long. Clearly the estimateprovided is exact only if f(x) is a linear function of x.The error made in calculating the area of the strip when the trapezium rule isused may be estimated as follows. The values used are f i <strong>and</strong> f i+1 , as in (27.36).These can be expressed accurately in terms of f i+1/2 <strong>and</strong> its derivatives by theTaylor seriesf i+1/2±1/2 = f i+1/2 ± h 2 f′ i+1/2 + 1 ( ) 2 hf i+1/2 ′′2! 2± 1 ( ) 3 hf (3)3! 2i+1/2 + ··· .ThusA i (estim.) = 1 2 h(f i + f i+1 )[= h f i+1/2 + 1 ( h2! 2whilst, from the first few terms of the exact result (27.35),A i (exact) = hf i+1/2 + 2 ( h3! 2Thus the error ∆A i = A i (estim.) − A i (exact) is given by∆A i = ( 18 − ) 124 h 3 f i+1/2 ′′ +O(h5 )) 2f ′′i+1/2 +O(h4 )) 3f ′′i+1/2 +O(h5 ).≈ 1 12 h3 f i+1/2 ′′ .The total error in I(estim.) is thus given approximately by],∆I(estim.) ≈ 1 12 nh3 〈f ′′ 〉 = 1 12 (b − a)h2 〈f ′′ 〉, (27.38)where 〈f ′′ 〉 represents an average value <strong>for</strong> the second derivative of f over theinterval a to b.◮Use the trapezium rule with h =0.5 to evaluateI =∫ 20(x 2 − 3x +4)dx,<strong>and</strong>, by evaluating the integral exactly, examine how well (27.38) estimates the error.With h =0.5, we will need five values of f(x) =x 2 − 3x + 4 <strong>for</strong> use in <strong>for</strong>mula (27.37).They are f(0) = 4, f(0.5) = 2.75, f(1) = 2, f(1.5) = 1.75 <strong>and</strong> f(2) = 2. Putting these into(27.37) givesI(estim.) = 0.5 (4 + 2 × 2.75 + 2 × 2 + 2× 1.75 + 2) = 4.75.2The exact value is[ ] x32I(exact) =3 − 3x22 +4x =4 2 . 301003


NUMERICAL METHODSThe difference between the estimate of the integral <strong>and</strong> the exact answer is 1/12. Equation(27.38) estimates this error as 2 × 0.25 ×〈f ′′ 〉/12. Our (deliberately chosen!) integr<strong>and</strong> isone <strong>for</strong> which 〈f ′′ 〉 can be evaluated trivially. Because f(x) is a quadratic function of x,its second derivative is constant, <strong>and</strong> equal to 2 in this case. Thus 〈f ′′ 〉 has value 2 <strong>and</strong>(27.38) estimates the error as 1/12; that the estimate is exactly right should be no surprisesince the Taylor expansion <strong>for</strong> a quadratic polynomial about any point always terminatesafter three terms <strong>and</strong> so no higher-order terms in h have been ignored in (27.38). ◭27.4.2 Simpson’s ruleWhereas the trapezium rule makes a linear interpolation of f, Simpson’s ruleeffectively mimics the local variation of f(x) using parabolas. The strips aretreated two at a time (figure 27.4(c)) <strong>and</strong> there<strong>for</strong>e their number, N, should bemade even.In the neighbourhood of x i ,<strong>for</strong>i odd, it is supposed that f(x) can be adequatelyrepresented by a quadratic <strong>for</strong>m,f(x i + y) =f i + ay + by 2 . (27.39)In particular, applying this to y = ±h yields two expressions involving bf i+1 = f(x i + h) =f i + ah + bh 2 ,f i−1 = f(x i − h) =f i − ah + bh 2 ;thusbh 2 = 1 2 (f i+1 + f i−1 − 2f i ).Now, in the representation (27.39), the area of the double strip from x i−1 tox i+1 is given byA i (estim.) =∫ h−h(f i + ay + by 2 ) dy =2hf i + 2 3 bh3 .Substituting <strong>for</strong> bh 2 then yields, <strong>for</strong> the estimated area,A i (estim.) = 2hf i + 2 3 h × 1 2 (f i+1 + f i−1 − 2f i )= 1 3 h(4f i + f i+1 + f i−1 ),an expression involving only given quantities. It should be noted that the valuesof neither b nor a need be calculated.For the full integral,(I(estim.) = 1 3 h f 0 + f N +4 ∑f m +2 ∑ )f m . (27.40)m oddm evenIt can be shown, by following the same procedure as in the trapezium rule case,that the error in the estimated area is approximately∆I(estim.) ≈(b − a)180 h4 〈f (4) 〉.1004


27.4 NUMERICAL INTEGRATION27.4.3 Gaussian integrationIn the cases considered in the previous two subsections, the function f wasmimicked by linear <strong>and</strong> quadratic functions. These yield exact answers if fitself is a linear or quadratic function (respectively) of x. This process couldbe continued by increasing the order of the polynomial mimicking-function soas to increase the accuracy with which more complicated functions f could benumerically integrated. However, the same effect can be achieved with less ef<strong>for</strong>tby not insisting upon equally spaced points x i .The detailed analysis of such methods of numerical integration, in which theintegration points are not equally spaced <strong>and</strong> the weightings given to the values ateach point do not fall into a few simple groups, is too long to be given in full here.Suffice it to say that the methods are based upon mimicking the given functionwith a weighted sum of mutually orthogonal polynomials. The polynomials, F n (x),are chosen to be orthogonal with respect to a particular weight function w(x), i.e.∫ baF n (x)F m (x)w(x) dx = k n δ nm ,where k n is some constant that may depend upon n. Often the weight function isunity <strong>and</strong> the polynomials are mutually orthogonal in the most straight<strong>for</strong>wardsense; this is the case <strong>for</strong> Gauss–Legendre integration <strong>for</strong> which the appropriatepolynomials are the Legendre polynomials, P n (x). This particular scheme isdiscussed in more detail below.Other schemes cover cases in which one or both of the integral limits a <strong>and</strong> bare not finite. For example, if the limits are 0 <strong>and</strong> ∞ <strong>and</strong> the integr<strong>and</strong> containsa negative exponential function e −αx , a simple change of variable can cast itinto a <strong>for</strong>m <strong>for</strong> which Gauss–Laguerre integration would be particularly wellsuited. This <strong>for</strong>m of quadrature is based upon the Laguerre polynomials, <strong>for</strong>which the appropriate weight function is w(x) =e −x . Advantage is taken of this,<strong>and</strong> the h<strong>and</strong>ling of the exponential factor in the integr<strong>and</strong> is effectively carriedout analytically. If the other factors in the integr<strong>and</strong> can be well mimicked bylow-order polynomials, then a Gauss–Laguerre integration using only a modestnumber of points gives accurate results.If we also add that the integral over the range −∞ to ∞ of an integr<strong>and</strong>containing an explicit factor exp(−βx 2 ) may be conveniently calculated using ascheme based on the Hermite polynomials, the reader will appreciate the closeconnection between the various Gaussian quadrature schemes <strong>and</strong> the sets ofeigenfunctions discussed in chapter 18. As noted above, the Gauss–Legendrescheme, which we discuss next, is just such a scheme, though its weight function,being unity throughout the range, is not explicitly displayed in the integr<strong>and</strong>.Gauss–Legendre quadrature can be applied to integrals over any finite rangethough the Legendre polynomials P l (x) on which it is based are only defined1005


NUMERICAL METHODS<strong>and</strong> orthogonal over the interval −1 ≤ x ≤ 1, as discussed in subsection 18.1.2.There<strong>for</strong>e, in order to use their properties, the integral between limits a <strong>and</strong> b in(27.34) has to be changed to one between the limits −1 <strong>and</strong> +1. This is easilydone with a change of variable from x to z given byz = 2x − b − a ,b − aso that I becomesI = b − a ∫ 1g(z) dz, (27.41)2 −1in which g(z) ≡ f(x).The n integration points x i <strong>for</strong> an n-point Gauss–Legendre integration aregiven by the zeros of P n (x), i.e. the x i are such that P n (x i ) = 0. The integr<strong>and</strong> g(x)is mimicked by the (n − 1)th-degree polynomialG(x) =n∑i=1P n (x)(x − x i )P ′ n(x i ) g(x i),which coincides with g(x) at each of the points x i , i =1, 2,...,n. To see this itshould be noted thatP n (x)limx→x k (x − x i )P n(x ′ i ) = δ ik.It then follows, to the extent that g(x) is well reproduced by G(x), that∫ 1−1g(x) dx ≈n∑i=1g(x i )P ′ n(x i )∫ 1−1P n (x)x − x idx. (27.42)The expressionw(x i ) ≡ 1 ∫ 1P n (x)P n(x ′ dxi ) −1 x − x ican be shown, using the properties of Legendre polynomials, to be equal to2w i =(1 − x 2 i )|P n(x ′ i )| , 2which is thus the weighting to be attached to the factor g(x i ) in the sum (27.42).The latter then becomes∫ 1n∑g(x) dx ≈ w i g(x i ). (27.43)−1In fact, because of the particular properties of Legendre polynomials, it can beshown that (27.43) integrates exactly any polynomial of degree up to 2n − 1. Theerror in the approximate equality is of the order of the 2nth derivative of g, <strong>and</strong>1006i=1


27.4 NUMERICAL INTEGRATIONso, provided g(x) is a reasonably smooth function, the approximation is a goodone.Taking 3-point integration as an example, the three x i are the zeros of P 3 (x) =12 (5x3 − 3x), namely 0 <strong>and</strong> ±0.774 60, <strong>and</strong> the corresponding weights are21 × ( − 2) 3 2= 8 2<strong>and</strong>9(1 − 0.6)× ( )6 2= 5 9 .2Table 27.8 gives the integration points (in the range −1 ≤ x i ≤ 1) <strong>and</strong> thecorresponding weights w i <strong>for</strong> a selection of n-point Gauss–Legendre schemes.◮Using a 3-point <strong>for</strong>mula in each case, evaluate the integral∫ 11I =0 1+x dx, 2(i) using the trapezium rule, (ii) using Simpson’s rule, (iii) using Gaussian integration. Alsoevaluate the integral analytically <strong>and</strong> compare the results.(i) Using the trapezium rule, we obtainI = 1 × [ ( 12 2 f(0) + 2f 1) ]2 + f(1)[= 1 4 1+8+ ] 15 2 =0.7750.(ii) Using Simpson’s rule, we obtainI = 1 × [ ( 13 2 f(0) + 4f 1) ]2 + f(1)[= 1 6 1+16+ ] 15 2 =0.7833.(iii) Using Gaussian integration, we obtainI = 1 − 0 ∫ 1dz2 −1 1+ 1 (z +1)24{}= 1 0.555 56 [f(−0.774 60) + f(0.774 60)] +0.888 89f(0)2{}= 1 0.555 56 [0.987 458 + 0.559 503] +0.888 89 × 0.82=0.785 27.(iv) Exact evaluation gives∫ 1dxI =0 1+x = [ tan −1 x ] 1= π =0.785 40.2 04In practice, a compromise has to be struck between the accuracy of the result achieved<strong>and</strong> the calculational labour that goes into obtaining it. ◭Further Gaussian quadrature procedures, ones that utilise the properties of theChebyshev polynomials, are available <strong>for</strong> integrals over finite ranges when theintegr<strong>and</strong>s involve factors of the <strong>for</strong>m (1 − x 2 ) ±1/2 . In the same way as decreasinglinear <strong>and</strong> quadratic exponentials are h<strong>and</strong>led through the weight functions inGauss–Laguerre <strong>and</strong> Gauss–Hermite quadrature, respectively, the square root1007


NUMERICAL METHODSGauss–Legendre integration∫ 1f(x) dx =−1i=1n∑w i f(x i )±x i w i ±x i w in =2 n =90.57735 02692 1.00000 00000 0.00000 00000 0.33023 935500.32425 34234 0.31234 70770n =3 0.61337 14327 0.26061 069640.00000 00000 0.88888 88889 0.83603 11073 0.18064 816070.77459 66692 0.55555 55556 0.96816 02395 0.08127 43884n =4 n =100.33998 10436 0.65214 51549 0.14887 43390 0.29552 422470.86113 63116 0.34785 48451 0.43339 53941 0.26926 671930.67940 95683 0.21908 63625n =5 0.86506 33667 0.14945 134920.00000 00000 0.56888 88889 0.97390 65285 0.06667 134430.53846 93101 0.47862 867050.90617 98459 0.23692 68851 n =120.12523 34085 0.24914 70458n =6 0.36783 14990 0.23349 253650.23861 91861 0.46791 39346 0.58731 79543 0.20316 742670.66120 93865 0.36076 15730 0.76990 26742 0.16007 832850.93246 95142 0.17132 44924 0.90411 72564 0.10693 932600.98156 06342 0.04717 53364n =70.00000 00000 0.41795 91837 n =200.40584 51514 0.38183 00505 0.07652 65211 0.15275 338710.74153 11856 0.27970 53915 0.22778 58511 0.14917 298650.94910 79123 0.12948 49662 0.37370 60887 0.14209 610930.51086 70020 0.13168 86384n =8 0.63605 36807 0.11819 453200.18343 46425 0.36268 37834 0.74633 19065 0.10193 011980.52553 24099 0.31370 66459 0.83911 69718 0.08327 674160.79666 64774 0.22238 10345 0.91223 44283 0.06267 204830.96028 98565 0.10122 85363 0.96397 19272 0.04060 142980.99312 85992 0.01761 40071Table 27.8 The integration points <strong>and</strong> weights <strong>for</strong> a number of n-pointGauss–Legendre integration <strong>for</strong>mulae. The points are given as ±x i <strong>and</strong> thecontributions from both +x i <strong>and</strong> −x i must be included. However, the contributionfrom any point x i = 0 must be counted only once.1008


27.4 NUMERICAL INTEGRATIONfactor is treated accurately in Gauss–Chebyshev integration. Thus∫ 1f(x)n∑√ dx ≈ w i f(x i ), (27.44)1 − x2−1where the integration points x i are the zeros of the Chebyshev polynomials ofthe first kind T n (x) <strong>and</strong>w i are the corresponding weights. Fortunately, both setsare analytic <strong>and</strong> can be written compactly <strong>for</strong> all n asx i =cos (i − 1 2 )π , w i = π <strong>for</strong> i =1,... ,n. (27.45)nnNote that, <strong>for</strong> any given n, all points are weighted equally <strong>and</strong> that no specialaction is required to deal with the integrable singularities at x = ±1; they aredealt with automatically through the weight function.For integrals involving factors of the <strong>for</strong>m (1−x 2 ) 1/2 , the corresponding <strong>for</strong>mula,based on Chebyshev polynomials of the second kind U n (x), is∫ 1f(x) √ n∑1 − x 2 dx ≈ w i f(x i ), (27.46)−1with integration points <strong>and</strong> weights given, <strong>for</strong> i =1,... ,n,byiπx i =cosn +1 , w i = π iπn +1 sin2 n +1 . (27.47)For discussions of the many other schemes available, as well as their relativemerits, the reader is referred to books devoted specifically to the theory ofnumerical analysis. There, details of integration points <strong>and</strong> weights, as well asquantitative estimates of the error involved in replacing an integral by a finitesum, will be found. Table 27.9 gives the points <strong>and</strong> weights <strong>for</strong> a selection ofGauss–Laguerre <strong>and</strong> Gauss–Hermite schemes. §i=1i=127.4.4 Monte Carlo methodsSurprising as it may at first seem, r<strong>and</strong>om numbers may be used to carry outnumerical integration. The r<strong>and</strong>om element comes in principally when selectingthe points at which the integr<strong>and</strong> is evaluated, <strong>and</strong> naturally does not extend tothe actual values of the integr<strong>and</strong>!For the most part we will continue to use as our model one-dimensionalintegrals between finite limits, as typified by equation (27.34). Extensions to coverinfinite or multidimensional integrals will be indicated briefly at the end of thesection. It should be noted here, however, that Monte Carlo methods – the name§ They, <strong>and</strong> those presented in table 27.8 <strong>for</strong> Gauss–Legendre integration, are taken from the muchmore comprehensive sets to be found in M. Abramowitz <strong>and</strong> I. A. Stegun (eds), H<strong>and</strong>book of<strong>Mathematical</strong> Functions (New York: Dover, 1965).1009


NUMERICAL METHODSGauss–Laguerre <strong>and</strong> Gauss–Hermite integration∫ ∞e −x f(x) dx =0i=1n∑w i f(x i )∫ ∞e −x2 f(x) dx =−∞i=1n∑w i f(x i )x i w i ±x i w in =2 n =20.58578 64376 0.85355 33906 0.70710 67812 0.88622 692553.41421 35624 0.14644 66094n =3n =3 0.00000 00000 1.18163 590060.41577 45568 0.71109 30099 1.22474 48714 0.29540 897522.29428 03603 0.27851 773366.28994 50829 0.01038 92565 n =40.52464 76233 0.80491 40900n =4 1.65068 01239 0.08131 283540.32254 76896 0.60315 410431.74576 11012 0.35741 86924 n =54.53662 02969 0.03888 79085 0.00000 00000 0.94530 872059.39507 09123 0.00053 92947 0.95857 24646 0.39361 932322.02018 28705 0.01995 32421n =50.26356 03197 0.52175 56106 n =61.41340 30591 0.39866 68111 0.43607 74119 0.72462 959523.59642 57710 0.07594 24497 1.33584 90740 0.15706 732037.08581 00059 0.00361 17587 2.35060 49737 0.00453 0009912.6408 00844 0.00002 33700n =7n =6 0.00000 00000 0.81026 461760.22284 66042 0.45896 46740 0.81628 78829 0.42560 725261.18893 21017 0.41700 08308 1.67355 16288 0.05451 558282.99273 63261 0.11337 33821 2.65196 13568 0.00097 178125.77514 35691 0.01039 919759.83746 74184 0.00026 10172 n =815.9828 73981 0.00000 08985 0.38118 69902 0.66114 701261.15719 37124 0.20780 23258n =7 1.98165 67567 0.01707 798300.19304 36766 0.40931 89517 2.93063 74203 0.00019 960411.02666 48953 0.42183 127792.56787 67450 0.14712 63487 n =94.90035 30845 0.02063 35145 0.00000 00000 0.72023 521568.18215 34446 0.00107 40101 0.72355 10188 0.43265 1559012.7341 80292 0.00001 58655 1.46855 32892 0.08847 4527419.3957 27862 0.00000 00317 2.26658 05845 0.00494 362433.19099 32018 0.00003 96070Table 27.9 The integration points <strong>and</strong> weights <strong>for</strong> a number of n-pointGauss–Laguerre <strong>and</strong> Gauss–Hermite integration <strong>for</strong>mulae. Where the pointsare given as ±x i , the contributions from both +x i <strong>and</strong> −x i must be included.However, the contribution from any point x i = 0 must be counted only once.1010


27.4 NUMERICAL INTEGRATIONhas become attached to methods based on r<strong>and</strong>omly generated numbers – inmany ways come into their own when used on multidimensional integrals overregions with complicated boundaries.It goes without saying that in order to use r<strong>and</strong>om numbers <strong>for</strong> calculationalpurposes a supply of them must be available. There was a time when they wereprovided in book <strong>for</strong>m as a two-dimensional array of r<strong>and</strong>om digits in the range0 to 9. The user could generate the successive digits of a r<strong>and</strong>om number ofany desired length by selecting their positions in the table in any predetermined<strong>and</strong> systematic way. Nowadays all computers <strong>and</strong> nearly all pocket calculatorsoffer a function which supplies a sequence of decimal numbers, ξ, that,<strong>for</strong>allpractical purposes, are r<strong>and</strong>omly <strong>and</strong> uni<strong>for</strong>mly chosen in the range 0 ≤ ξ


NUMERICAL METHODSaveraged:t = 1 nn∑f(ξ i ). (27.49)i=1Stratified samplingHere the range of x is broken up into k subranges,0=α 0


27.4 NUMERICAL INTEGRATIONwill have a very small variance. Further, any error in inverting the relationshipbetween η <strong>and</strong> ξ will not be important since f(η)/g(η) will be largely independentof the value of η.As an example, consider the function f(x) =[tan −1 (x)] 1/2 , which is not analyticallyintegrable over the range (0, 1) but is well mimicked by the easily integratedfunction g(x) =x 1/2 (1 − x 2 /6). The ratio of the two varies from 1.00 to 1.06 as xvaries from 0 to 1. The integral of g over this range is 0.619 048, <strong>and</strong> so it has tobe renormalised by the factor 1.615 38. The value of the integral of f(x) from0to 1 can then be estimated by averaging the value of[tan −1 (η)] 1/2(1.615 38) η 1/2 (1 − 1 6 η2 )<strong>for</strong> r<strong>and</strong>om variables η which are such that G(η) is uni<strong>for</strong>mly distributed on(0, 1). Using batches of as few as ten r<strong>and</strong>om numbers gave a value 0.630 <strong>for</strong> θ,with st<strong>and</strong>ard deviation 0.003. The corresponding result <strong>for</strong> crude Monte Carlo,using the same r<strong>and</strong>om numbers, was 0.634 ± 0.065. The increase in precision isobvious, though the additional labour involved would not be justified <strong>for</strong> a singleapplication.Control variatesThe control-variate method is similar to, but not the same as, importance sampling.Again, an analytically integrable function that mimics f(x) in shape hasto be found. The function, known as the control variate, is first scaled so as tomatch f as closely as possible in magnitude <strong>and</strong> then its integral is found inclosed <strong>for</strong>m. If we denote the scaled control variate by h(x), then the estimate ofθ is computed ast =∫ 10[f(x) − h(x)] dx +∫ 10h(x) dx. (27.51)The first integral in (27.51) is evaluated using (crude) Monte Carlo, whilst thesecond is known analytically. Although the first integral should have been renderedsmall by the choice of h(x), it is its variance that matters. The method relieson the following result (see equation (30.136)):V [t − t ′ ]=V [t]+V [t ′ ] − 2Cov[t, t ′ ],<strong>and</strong> on the fact that if t estimates θ whilst t ′ estimates θ ′ using the same r<strong>and</strong>omnumbers, then the covariance of t <strong>and</strong> t ′ can be larger than the variance of t ′ ,<strong>and</strong> indeed will be so if the integr<strong>and</strong>s producing θ <strong>and</strong> θ ′ are highly correlated.To evaluate the same integral as was estimated previously using importancesampling, we take as h(x) the function g(x) used there, be<strong>for</strong>e it was renormalised.Again using batches of ten r<strong>and</strong>om numbers, the estimated value <strong>for</strong> θ was foundto be 0.629 ± 0.004, a result almost identical to that obtained using importance1013


NUMERICAL METHODSsampling, in both value <strong>and</strong> precision. Since we knew already that f(x) <strong>and</strong>g(x)diverge monotonically by about 6% as x varies over the range (0, 1), we couldhave made a small improvement to our control variate by scaling it by 1.03 be<strong>for</strong>eusing it in equation (27.51).Antithetic variatesAs a final example of a method that improves on crude Monte Carlo, <strong>and</strong> one thatis particularly useful when monotonic functions are to be integrated, we mentionthe use of antithetic variates. This method relies on finding two estimates t <strong>and</strong>t ′ of θ that are strongly anticorrelated (i.e. Cov[t, t ′ ] is large <strong>and</strong> negative) <strong>and</strong>using the resultV [ 1 2 (t + t′ )] = 1 4 V [t]+ 1 4 V [t′ ]+ 1 2 Cov[t, t′ ].For example, the use of 1 2[f(ξ) +f(1 − ξ)] instead of f(ξ) involves only twiceas many evaluations of f, <strong>and</strong> no more r<strong>and</strong>om variables, but generally givesan improvement in precision significantly greater than this. For the integral off(x) =[tan −1 (x)] 1/2 , using as previously a batch of ten r<strong>and</strong>om variables, anestimate of 0.623 ± 0.018 was found. This is to be compared with the crudeMonte Carlo result, 0.634 ± 0.065, obtained using the same number of r<strong>and</strong>omvariables.For a fuller discussion of these methods, <strong>and</strong> of theoretical estimates of theirefficiencies, the reader is referred to more specialist treatments. For practical implementationschemes, a book dedicated to scientific computing should be consulted. §Hit or miss methodWe now come to the approach that, in spirit, is closest to the activities that gaveMonte Carlo methods their name. In this approach, one or more straight<strong>for</strong>wardyes/no decisions are made on the basis of numbers drawn at r<strong>and</strong>om – the endresult of each trial is either a hit or a miss! In this section we are concernedwith numerical integration, but the general Monte Carlo approach, in whichone estimates a physical quantity that is hard or impossible to calculate directlyby simulating the physical processes that determine it, is widespread in modernscience. For example, the calculation of the efficiencies of detector arrays inexperiments to study elementary particle interactions are nearly always carriedout in this way. Indeed, in a normal experiment, far more simulated interactionsare generated in computers than ever actually occur when the experiment istaking real data.As was noted in chapter 2, the process of evaluating a one-dimensional integralf(x)dx can be regarded as that of finding the area between the curve y = f(x)∫ ba§ e.g. W. H. Press, S. A. Teukolsky, W. T. Vetterling <strong>and</strong> B. P. Flannery, Numerical Recipes in C: TheArtofScientificComputing, 2nd edn (Cambridge: Cambridge University Press, 1992).1014


27.4 NUMERICAL INTEGRATIONy = f(x)y = cx = ax = bxFigure 27.5 A simple rectangular figure enclosing the area (shown shaded)which is equal to ∫ bf(x) dx.a<strong>and</strong> the x-axis in the range a ≤ x ≤ b. It may not be possible to do thisanalytically, but if, as shown in figure 27.5, we can enclose the curve in a simplefigure whose area can be found trivially then the ratio of the required (shaded)area to that of the bounding figure, c(b − a), is the same as the probability that ar<strong>and</strong>omly selected point inside the boundary will lie below the line.In order to accommodate cases in which f(x) can be negative in part of thex-range, we treat a slightly more general case. Suppose that, <strong>for</strong> a ≤ x ≤ b, f(x)is bounded <strong>and</strong> known to lie in the range A ≤ f(x) ≤ B; then the trans<strong>for</strong>mationz = x − ab − awill reduce the integral ∫ baf(x) dx to the <strong>for</strong>mA(b − a)+(B − A)(b − a)∫ 10h(z) dz, (27.52)whereh(z) = 1 [f ((b − a)z + a) − A] .B − AIn this <strong>for</strong>m z lies in the range 0 ≤ z ≤ 1<strong>and</strong>h(z) lies in the range 0 ≤ h(z) ≤ 1,i.e. both are suitable <strong>for</strong> simulation using the st<strong>and</strong>ard r<strong>and</strong>om-number generator.It should be noted that, <strong>for</strong> an efficient estimation, the bounds A <strong>and</strong> B shouldbe drawn as tightly as possible –preferably, but not necessarily, they should beequal to the minimum <strong>and</strong> maximum values of f in the range. The reason <strong>for</strong>this is that r<strong>and</strong>om numbers corresponding to values which f(x) cannot reachadd nothing to the estimation but do increase its variance.It only remains to estimate the final integral on the RHS of equation (27.52).This we do by selecting pairs of r<strong>and</strong>om numbers, ξ 1 <strong>and</strong> ξ 2 , <strong>and</strong> testing whether1015


NUMERICAL METHODSh(ξ 1 ) >ξ 2 . The fraction of times that this inequality is satisfied estimates the valueof the integral (without the scaling factors (B − A)(b − a)) since the expectationvalue of this fraction is the ratio of the area below the curve y = h(z) totheareaof a unit square.To illustrate the evaluation of multiple integrals using Monte Carlo techniques,consider the relatively elementary problem of finding the volume of an irregularsolid bounded by planes, say an octahedron. In order to keep the descriptionbrief, but at the same time illustrate the general principles involved, let us supposethat the octahedron has two vertices on each of the three Cartesian axes, one oneither side of the origin <strong>for</strong> each axis. Denote those on the x-axis by x 1 (< 0) <strong>and</strong>x 2 (> 0), <strong>and</strong> similarly <strong>for</strong> the y- <strong>and</strong>z-axes. Then the whole of the octahedroncan be enclosed by the rectangular parallelepipedx 1 ≤ x ≤ x 2 , y 1 ≤ y ≤ y 2 , z 1 ≤ z ≤ z 2 .Any point in the octahedron lies inside or on the parallelepiped, but any pointin the parallelepiped may or may not lie inside the octahedron.The equation of the plane containing the three vertex points (x i , 0, 0), (0,y j , 0)<strong>and</strong> (0, 0,z k )isx+ y + z =1 <strong>for</strong>i, j, k =1, 2, (27.53)x i y j z k<strong>and</strong> the condition that any general point (x, y, z) lies on the same side of theplane as the origin is thatx+ y + z − 1 ≤ 0. (27.54)x i y j z kFor the point to be inside or on the octahedron, equation (27.54) must there<strong>for</strong>ebe satisfied <strong>for</strong> all eight of the sets of i, j <strong>and</strong> k given in (27.53).Thus an estimate of the volume of the octahedron can be made by generatingr<strong>and</strong>om numbers ξ from the usual uni<strong>for</strong>m distribution <strong>and</strong> then using them insets of three, according to the following scheme.With integer m labelling the mth set of three r<strong>and</strong>om numbers, calculatex = x 1 + ξ 3m−2 (x 2 − x 1 ),y = y 1 + ξ 3m−1 (y 2 − y 1 ),z = z 1 + ξ 3m (z 2 − z 1 ).Define a variable n m as 1 if (27.54) is satisfied <strong>for</strong> all eight combinations of i, j, kvalues <strong>and</strong> as 0 otherwise. The volume V can then be estimated using 3M r<strong>and</strong>omnumbers from the <strong>for</strong>mulaV(x 2 − x 1 )(y 2 − y 1 )(z 2 − z 1 ) = 1 M1016M∑n m .m=1


27.4 NUMERICAL INTEGRATIONIt will be seen that, by replacing each n m in the summation by f(x, y, z)n m ,this procedure could be extended to estimate the integral of the function f overthe volume of the solid. The method has special valueif f is too complicated tohave analytic integrals with respect to x, y <strong>and</strong> z or if the limits of any of theseintegrals are determined by anything other than the simplest combinations of theother variables. If large values of f are known to be concentrated in particularregions of the integration volume, then some <strong>for</strong>m of stratified sampling shouldbe used.It will be apparent that this general method can be extended to integralsof general functions, bounded but not necessarily continuous, over volumes withcomplicated bounding surfaces <strong>and</strong>, if appropriate, in more than three dimensions.R<strong>and</strong>om number generationEarlier in this subsection we showed how to evaluate integrals using sequences ofnumbers that we took to be distributed uni<strong>for</strong>mly on the interval 0 ≤ ξ


NUMERICAL METHODSon (0, 1) <strong>and</strong> then take as the r<strong>and</strong>om number y the value of F −1 (ξ). We nowillustrate this with a worked example.◮Find an explicit <strong>for</strong>mula that will generate a r<strong>and</strong>om number y distributed on (−∞, ∞)according to the Cauchy distribution( a)dyf(y) dy =π a 2 + y , 2given a r<strong>and</strong>om number ξ uni<strong>for</strong>mly distributed on (0, 1).The first task is to determine the indefinite integral:∫ y ( a)dtF(y) =−∞ π a 2 + t = 1 y 2 π tan−1 a + 1 2 .Now, if y is distributed as we wish then F(y) is uni<strong>for</strong>mly distributed on (0, 1). This followsfrom the fact that the derivative of F(y) isf(y). We there<strong>for</strong>e set F(y) equaltoξ <strong>and</strong>obtainξ = 1 y π tan−1 a + 1 2 ,yieldingy = a tan[π(ξ − 1 )]. 2This explicit <strong>for</strong>mula shows how to change a r<strong>and</strong>om number ξ drawn from a populationuni<strong>for</strong>mly distributed on (0, 1) into a r<strong>and</strong>om number y distributed according to theCauchy distribution. ◭Look-up tables operate as described below <strong>for</strong> cumulative distributions F(y)that are non-invertible, i.e. F −1 (y) cannot be expressed in closed <strong>for</strong>m. Theyare especially useful if many r<strong>and</strong>om numbers are needed but great samplingaccuracy is not essential. The method <strong>for</strong> an N-entry table can be summarised asfollows. Define w m by F(w m )=m/N <strong>for</strong> m =1, 2,...,N, <strong>and</strong> store a table ofy(m) = 1 2 (w m + w m−1 ).As each r<strong>and</strong>om number y is needed, calculate k as the integral part of Nξ <strong>and</strong>take y as given by y(k).Normally, such a look-up table would have to be used <strong>for</strong> generating r<strong>and</strong>omnumbers with a Gaussian distribution, as the cumulative integral of a Gaussian isnon-invertible. It would be, in essence, table 30.3, with the roles of argument <strong>and</strong>value interchanged. In this particular case, an alternative, based on the centrallimit theorem, can be considered.With ξ i generated in the usual way, i.e. uni<strong>for</strong>mly distributed on the interval0 ≤ ξ


27.5 FINITE DIFFERENCESmany values of ξ i <strong>for</strong> each value of y <strong>and</strong> is a very poor approximation if thewings of the Gaussian distribution have to be sampled accurately. For nearly allpractical purposes a Gaussian look-up table is to be preferred.27.5 Finite differencesIt will have been noticed that earlier sections included several equations linkingsequential values of f i <strong>and</strong> the derivatives of f evaluated at one of the x i .Inthis section, by way of preparation <strong>for</strong> the numerical treatment of differentialequations, we establish these relationships in a more systematic way.Again we consider a set of values f i of a function f(x) evaluated at equallyspaced points x i , their separation being h. As be<strong>for</strong>e, the basis <strong>for</strong> our discussionwill be a Taylor series expansion, but on this occasion about the point x i :f i±1 = f i ± hf i ′ + h22! f′′ i ± h33! f(3) i + ··· . (27.56)In this section, <strong>and</strong> subsequently, we denote the nth derivative evaluated at x iby f (n)i .From (27.56), three different expressions that approximate f (1)ican be derived.The first of these, obtained by subtracting the ± equations, is( )f (1) dfi≡ = f i+1 − f i−1dx 2hx i− h23! f(3) i− ··· . (27.57)The quantity (f i+1 − f i−1 )/(2h) is known as the central difference approximationto f (1)i<strong>and</strong> can be seen from (27.57) to be in error by approximately (h 2 /6)f (3)i.An alternative approximation, obtained from (27.56+) alone, is given by( )f (1) dfi ≡ = f i+1 − f i− h dxx ih 2! f(2) i − ··· . (27.58)The <strong>for</strong>ward difference approximation, (f i+1 − f i )/h, is clearly a poorer approximation,since it is in error by approximately (h/2)f (2)i as compared with (h 2 /6)f (3)i .Similarly, the backward difference (f i − f i−1 )/h obtained from (27.56−) is not asgood as the central difference; the sign of the error is reversed in this case.This type of differencing approximation can be continued to the higher derivativesof f in an obvious manner. By adding the two equations (27.56±), a centraldifference approximation to f (2)i can be obtained:(f (2) d 2 )fi≡dx 2 ≈ f i+1 − 2f i + f i−1h 2 . (27.59)The error in this approximation (also known as the second difference of f) iseasilyshowntobeabout(h 2 /12)f (4)i.Of course, if the function f(x) is a sufficiently simple polynomial in x, all1019


NUMERICAL METHODSderivatives beyond a particular one will vanish <strong>and</strong> there is no error in takingthe differences to obtain the derivatives.◮The following is copied from the tabulation of a second-degree polynomial f(x) at valuesof x from 1 to 12 inclusive:2, 2, ?, 8, 14, 22, 32, 46, ?, 74, 92, 112.The entries marked ? were illegible <strong>and</strong> in addition one error was made in transcription.Complete <strong>and</strong> correct the table. Would your procedure have worked if the copying errorhad been in f(6)?Write out the entries again in row (a) below, <strong>and</strong> where possible calculate first differencesin row (b) <strong>and</strong> second differences in row (c). Denote the jth entry in row (n) by(n) j .(a) 2 2 ? 8 14 22 32 46 ? 74 92 112(b) 0 ? ? 6 8 10 14 ? ? 18 20(c) ? ? ? 2 2 4 ? ? ? 2Because the polynomial is second-degree, the second differences (c) j , which are proportionalto d 2 f/dx 2 , should be constant, <strong>and</strong> clearly the constant should be 2. That is, (c) 6 shouldequal 2 <strong>and</strong> (b) 7 should equal 12 (not 14). Since all the (c) j = 2, we can conclude that(b) 2 =2,(b) 3 =4,(b) 8 = 14, <strong>and</strong> (b) 9 = 16. Working these changes back to row (a) showsthat (a) 3 =4,(a) 8 = 44 (not 46), <strong>and</strong> (a) 9 = 58.The entries there<strong>for</strong>e should read(a) 2, 2, 4, 8, 14, 22, 32, 44, 58, 74, 92, 112,where the amended entries are shown in bold type.It is easily verified that if the error were in f(6) no two computable entries in row (c)would be equal, <strong>and</strong> it would not be clear what the correct common entry should be.Nevertheless, trial <strong>and</strong> error might arrive at a self-consistent scheme. ◭27.6 Differential equationsFor the remaining sections of this chapter our attention will be on the solutionof differential equations by numerical methods. Some of the general difficultiesof applying numerical methods to differential equations will be all too apparent.Initially we consider only the simplest kind of equation – one of first order,typically represented bydy= f(x, y), (27.60)dxwhere y is taken as the dependent variable <strong>and</strong> x the independent one. If thisequation can be solved analytically then that is the best course to adopt. Butsometimes it is not possible to do so <strong>and</strong> a numerical approach becomes theonly one available. In fact, most of the examples that we will use can be solvedeasily by an explicit integration, but, <strong>for</strong> the purposes of illustration, this is anadvantage rather than the reverse since useful comparisons can then be madebetween the numerically derived solution <strong>and</strong> the exact one.1020


27.6 DIFFERENTIAL EQUATIONSx h y(exact)0.01 0.1 0.5 1.0 1.5 2 30 (1) (1) (1) (1) (1) (1) (1) (1)0.5 0.605 0.590 0.500 0 −0.500 −1 −2 0.6071.0 0.366 0.349 0.250 0 0.250 1 4 0.3681.5 0.221 0.206 0.125 0 −0.125 −1 −8 0.2232.0 0.134 0.122 0.063 0 0.063 1 16 0.1352.5 0.081 0.072 0.032 0 −0.032 −1 −32 0.0823.0 0.049 0.042 0.016 0 0.016 1 64 0.050Table 27.10 The solution y of differential equation (27.61) using the Euler<strong>for</strong>ward difference method <strong>for</strong> various values of h. The exact solution is alsoshown.27.6.1 Difference equationsConsider the differential equationdy= −y, y(0) = 1, (27.61)dx<strong>and</strong> the possibility of solving it numerically by approximating dy/dx by a finitedifference along the lines indicated in section 27.5. We start with the <strong>for</strong>warddifference( ) dy≈ y i+1 − y i, (27.62)dxx ihwhere we use the notation of section 27.5 but with f replaced by y. Inthisparticular case, it leads to the recurrence relation( ) dyy i+1 = y i + h = y i − hy i =(1− h)y i . (27.63)dxiThus, since y 0 = y(0) = 1 is given, y 1 = y(0 + h) =y(h) can be calculated, <strong>and</strong>so on (this is the Euler method). Table 27.10 shows the values of y(x) obtainedif this is done using various values of h <strong>and</strong> <strong>for</strong> selected values of x. The exactsolution, y(x) =exp(−x), is also shown.It is clear that to maintain anything like a reasonable accuracy only very smallsteps, h, can be used. Indeed, if h is taken to be too large, not only is the accuracybad but, as can be seen, <strong>for</strong> h>1 the calculated solution oscillates (when itshould be monotonic), <strong>and</strong> <strong>for</strong> h>2 it diverges. Equation (27.63) is of the <strong>for</strong>my i+1 = λy i , <strong>and</strong> a necessary condition <strong>for</strong> non-divergence is |λ| < 1, i.e. 0


NUMERICAL METHODSx y(estim.) y(exact)−0.5 (1.648) —0 (1.000) (1.000)0.5 0.648 0.6071.0 0.352 0.3681.5 0.296 0.2232.0 0.056 0.1352.5 0.240 0.0823.0 −0.184 0.050Table 27.11 The solution of differential equation (27.61) using the Milnecentral difference method with h =0.5 <strong>and</strong> accurate starting values.more accurate, but of course still approximate, central difference. A more accuratemethod based on central differences (Milne’s method) gives the recurrence relation( ) dyy i+1 = y i−1 +2h(27.64)dxiin general <strong>and</strong>, in this particular case,y i+1 = y i−1 − 2hy i . (27.65)An additional difficulty now arises, since two initial values of y are needed.The second must be estimated by other means (e.g. by using a Taylor series,as discussed later), but <strong>for</strong> illustration purposes we will take the accurate value,y(−h) =exph, as the value of y −1 .Ifh is taken as, say, 0.5 <strong>and</strong> (27.65) is appliedrepeatedly, then the results shown in table 27.11 are obtained.Although some improvement in the early values of the calculated y(x) isnoticeable, as compared with the corresponding (h =0.5) column of table 27.10,this scheme soon runs into difficulties, as is obvious from the last two rows of thetable.Some part of this poor per<strong>for</strong>mance is not really attributable to the approximationsmade in estimating dy/dx but to the <strong>for</strong>m of the equation itself <strong>and</strong>hence of its solution. Any rounding error occurring in the evaluation effectivelyintroduces into y some contamination by the solution ofdydx =+y.This equation has the solution y(x) =expx <strong>and</strong> so grows without limit; ultimatelyit will dominate the sought-<strong>for</strong> solution <strong>and</strong> thus render the calculations totallyinaccurate.We have only illustrated, rather than analysed, some of the difficulties associatedwith simple finite-difference iteration schemes <strong>for</strong> first-order differential equations,1022


27.6 DIFFERENTIAL EQUATIONSbut they may be summarised as (i) insufficiently precise approximations to thederivatives <strong>and</strong> (ii) inherent instability due to rounding errors.27.6.2 Taylor series solutionsSince a Taylor series expansion is exact if all its terms are included, <strong>and</strong> the limitsof convergence are not exceeded, we may seek to use one to evaluate y 1 , y 2 ,etc.<strong>for</strong> an equationdy= f(x, y), (27.66)dxwhen the initial value y(x 0 )=y 0 is given.The Taylor series isy(x + h) =y(x)+hy ′ (x)+ h22! y′′ (x)+ h33! y(3) (x)+··· . (27.67)In the present notation, at the point x = x i this is writteny i+1 = y i + hy (1)i + h22! y(2) i+ h33! y(3) i + ··· . (27.68)But, <strong>for</strong> the required solution y(x), we know that( )y (1) dyi ≡ = f(x i ,y i ), (27.69)dxx i<strong>and</strong> the value of the second derivative at x = x i , y = y i can be obtained from it:y (2)i= ∂f∂x + ∂f dy∂y dx = ∂f∂x + f ∂f∂y . (27.70)This process can be continued <strong>for</strong> the third <strong>and</strong> higher derivatives, all of whichare to be evaluated at (x i ,y i ).Having obtained expressions <strong>for</strong> the derivatives y (n)iin (27.67), two alternativeways of proceeding are open to us:(i) equation (27.68) is used to evaluate y i+1 , the whole process is repeated toobtain y i+2 ,<strong>and</strong>soon;(ii) equation (27.68) is applied several times but using a different value of heach time, <strong>and</strong> so the corresponding values of y(x + h) are obtained.It is clear that, on the one h<strong>and</strong>, approach (i) does not require so many terms of(27.67) to be kept, but, on the other h<strong>and</strong>, the y i (n) have to be recalculated ateach step. With approach (ii), fairly accurate results <strong>for</strong> y may be obtained <strong>for</strong>values of x close to the given starting value, but <strong>for</strong> large values of h a largenumber of terms of (27.67) must be kept. As an example of approach (ii) wesolve the following problem.1023


NUMERICAL METHODSx y(estim.) y(exact)0 1.0000 1.00000.1 1.2346 1.23460.2 1.5619 1.56250.3 2.0331 2.04080.4 2.7254 2.77780.5 3.7500 4.0000Table 27.12The solution of differential equation (27.71) using a Taylor series.◮Find the numerical solution of the equationdydx =2y3/2 , y(0) = 1, (27.71)<strong>for</strong> x =0.1 to 0.5 in steps of 0.1. Compare it with the exact solution obtained analytically.Since the right-h<strong>and</strong> side of the equation does not contain x explicitly, (27.70) is greatlysimplified <strong>and</strong> the calculation becomes a repeated application ofy (n+1)i= ∂y(n) dy∂y dx = f ∂y(n)∂y .The necessary derivatives <strong>and</strong> their values at x =0,wherey = 1, are given below:y(0) = 1 1y ′ =2y 3/2 2y ′′ =(3/2)(2y 1/2 )(2y 3/2 )=6y 2 6y (3) =(12y)2y 3/2 =24y 5/2 24y (4) =(60y 3/2 )2y 3/2 = 120y 3 120y (5) = (360y 2 )2y 3/2 = 720y 7/2 720Thus the Taylor expansion of the solution about the origin (in fact a Maclaurin series) isy(x)=1+2x + 6 2! x2 + 243! x3 + 1204! x4 + 7205! x5 + ··· .Hence, y(estim.) = 1 + 2x +3x 2 +4x 3 +5x 4 +6x 5 . Values calculated from this are givenin table 27.12. Comparison with the exact values shows that using the first six terms givesa value that is correct to one part in 100, up to x =0.3. ◭27.6.3 Prediction <strong>and</strong> correctionAn improvement in the accuracy obtainable using difference methods is possibleif steps are taken, sometimes retrospectively, to allow <strong>for</strong> inaccuracies in approximatingderivatives by differences. We will describe only the simplest schemes ofthis kind <strong>and</strong> begin with a prediction method, usually called the Adams method.1024


27.6 DIFFERENTIAL EQUATIONSThe <strong>for</strong>ward difference estimate of y i+1 , namely( ) dyy i+1 = y i + h = y i + hf(x i ,y i ), (27.72)dxiwould give exact results if y were a linear function of x in the range x i ≤ x ≤ x i +h.The idea behind the Adams method is to allow some relaxation of this <strong>and</strong>suppose that y can be adequately approximated by a parabola over the intervalx i−1 ≤ x ≤ x i+1 . In the same interval, dy/dx can then be approximated by a linearfunction:f(x, y) = dydx ≈ a + b(x − x i) <strong>for</strong> x i − h ≤ x ≤ x i + h.The values of a <strong>and</strong> b are fixed by the calculated values of f at x i−1 <strong>and</strong> x i ,whichwe may denote by f i−1 <strong>and</strong> f i :Thuswhich yieldsa = f i ,∫ xi+h[y i+1 − y i ≈x ib = f i − f i−1.hf i + (f i − f i−1 )h](x − x i ) dx,y i+1 = y i + hf i + 1 2 h(f i − f i−1 ). (27.73)The last term of this expression is seen to be a correction to result (27.72). Thatit is, in some sense, the second-order correction,12 h2 y (2)i−1/2 ,to a first-order <strong>for</strong>mula is apparent.Such a procedure requires, in addition to a value <strong>for</strong> y 0 , a value <strong>for</strong> either y 1 ory −1 ,sothatf 1 or f −1 can be used to initiate the iteration. This has to be obtainedby other methods, e.g. a Taylor series expansion.Improvements to simple difference <strong>for</strong>mulae can also be obtained by usingcorrection methods. In these, a rough prediction of the value y i+1 is made first,<strong>and</strong> then this is used in a better <strong>for</strong>mula, not originally usable since it, in turn,requires a value of y i+1 <strong>for</strong> its evaluation. The value of y i+1 is then recalculated,using this better <strong>for</strong>mula.Such a scheme based on the <strong>for</strong>ward difference <strong>for</strong>mula might be as follows:(i) predict y i+1 using y i+1 = y i + hf i ;(ii) calculate f i+1 using this value;(iii) recalculate y i+1 using y i+1 = y i + h(f i + f i+1 )/2. Here (f i + f i+1 )/2 hasreplaced the f i used in (i), since it better represents the average value ofdy/dx in the interval x i ≤ x ≤ x i + h.1025


NUMERICAL METHODSSteps (ii) <strong>and</strong> (iii) can be iterated to improve further the approximation to theaverage value of dy/dx, but this will not compensate <strong>for</strong> the omission of higherorderderivatives in the <strong>for</strong>ward difference <strong>for</strong>mula.Many more complex schemes of prediction <strong>and</strong> correction, in most casescombining the two in the same process, have been devised, but the reader isreferred to more specialist texts <strong>for</strong> discussions of them. However, because itoffers some clear advantages, one group of methods will be set out explicitly inthe next subsection. This is the general class of schemes known as Runge–Kuttamethods.27.6.4 Runge–Kutta methodsThe Runge–Kutta method of integratingdy= f(x, y) (27.74)dxis a step-by-step process of obtaining an approximation <strong>for</strong> y i+1 by starting fromthe value of y i . Among its advantages are that no functions other than f are used,no subsidiary differentiation is needed <strong>and</strong> no additional starting values need becalculated.To be set against these advantages is the fact that f is evaluated using somewhatcomplicated arguments <strong>and</strong> that this has to be done several times <strong>for</strong> each increasein the value of i. However, once a procedure has been established, <strong>for</strong> exampleon a computer, the method usually gives good results.The basis of the method is to simulate the (accurate) Taylor series <strong>for</strong> y(x i + h),not by calculating all the higher derivatives of y at the point x i but by takinga particular combination of the values of the first derivative of y evaluated ata number of carefully chosen points. Equation (27.74) is used to evaluate thesederivatives. The accuracy can be made to be up to whatever power of h is desired,but, naturally, the greater the accuracy, the more complex the calculation, <strong>and</strong>,in any case, rounding errors cannot ultimately be avoided.The setting up of the calculational scheme may be illustrated by consideringthe particular case in which second-order accuracy in h is required. To secondorder, the Taylor expansion is( )y i+1 = y i + hf i + h2 df, (27.75)2 dxx iwhere( ) ( df ∂f=dxx i∂x + f ∂f )≡ ∂f i∂yx i∂x + f ∂f ii∂y ,the last step being merely the definition of an abbreviated notation.1026


27.6 DIFFERENTIAL EQUATIONSWe assume that this can be simulated by a <strong>for</strong>my i+1 = y i + α 1 hf i + α 2 hf(x i + β 1 h, y i + β 2 hf i ), (27.76)which in effect uses a weighted mean of the value of dy/dx at x i <strong>and</strong> its value atsome point yet to be determined. The object is to choose values of α 1 , α 2 , β 1 <strong>and</strong>β 2 such that (27.76) coincides with (27.75) up to the coefficient of h 2 .Exp<strong>and</strong>ing the function f in the last term of (27.76) in a Taylor series of itsown, we obtainf(x i + β 1 h, y i + β 2 hf i )=f(x i ,y i )+β 1 h ∂f i∂x + β 2hf i∂f i∂y +O(h2 ).Putting this result into (27.76) <strong>and</strong> rearranging in powers of h, we obtain()y i+1 = y i +(α 1 + α 2 )hf i + α 2 h 2 ∂f iβ 1∂x + β ∂f i2f i . (27.77)∂yComparing this with (27.75) shows that there is, in fact, some freedom remainingin the choice of the α’s <strong>and</strong> β’s. In terms of an arbitrary α 1 (≠1),α 2 =1− α 1 , β 1 = β 2 =12(1 − α 1 ) .One possible choice is α 1 =0.5, giving α 2 =0.5, β 1 = β 2 = 1. In this case theprocedure (equation (27.76)) can be summarised bywherey i+1 = y i + 1 2 (a 1 + a 2 ), (27.78)a 1 = hf(x i ,y i ),a 2 = hf(x i + h, y i + a 1 ).Similar schemes giving higher-order accuracy in h can be devised. Two suchschemes, given without derivation, are as follows.(i) To order h 3 ,wherey i+1 = y i + 1 6 (b 1 +4b 2 + b 3 ), (27.79)b 1 = hf(x i ,y i ),b 2 = hf(x i + 1 2 h, y i + 1 2 b 1),b 3 = hf(x i + h, y i +2b 2 − b 1 ).1027


NUMERICAL METHODS(ii) To order h 4 ,wherey i+1 = y i + 1 6 (c 1 +2c 2 +2c 3 + c 4 ), (27.80)c 1 = hf(x i ,y i ),c 2 = hf(x i + 1 2 h, y i + 1 2 c 1),c 3 = hf(x i + 1 2 h, y i + 1 2 c 2),c 4 = hf(x i + h, y i + c 3 ).27.6.5 IsoclinesThe final method to be described <strong>for</strong> first-order differential equations is not somuch numerical as graphical, but since it is sometimes useful it is included here.The method, known as that of isoclines, involves sketching <strong>for</strong> a number ofvalues of a parameter c those curves (the isoclines) in the xy-plane along whichf(x, y) =c, i.e. those curves along which dy/dx is a constant of known value. Itshould be noted that isoclines are not generally straight lines. Since a straightline of slope dy/dx at <strong>and</strong> through any particular point is a tangent to the curvey = y(x) at that point, small elements of straight lines, with slopes appropriateto the isoclines they cut, effectively <strong>for</strong>m the curve y = y(x).Figure 27.6 illustrates in outline the method as applied to the solution ofdy= −2xy. (27.81)dxThe thinner curves (rectangular hyperbolae) are a selection of the isoclines alongwhich −2xy is constant <strong>and</strong> equal to the corresponding value of c. The smallcross lines on each curve show the slopes (= c) that solutions of (27.81) musthave if they cross the curve. The thick line is the solution <strong>for</strong> which y =1atx = 0; it takes the slope dictated by the value of c on each isocline it crosses. Theanalytic solution with these properties is y(x) =exp(−x 2 ).27.7 Higher-order equationsSo far the discussion of numerical solutions of differential equations has beenin terms of one dependent <strong>and</strong> one independent variable related by a first-orderequation. It is straight<strong>for</strong>ward to carry out an extension to the case of severaldependent variables y [r] governed by R first-order equations:dy [r]dx = f [r](x, y [1] ,y [2] ,...,y [R] ), r =1, 2,...,R.We have enclosed the label r in brackets so that there is no confusion between,say, the second dependent variable y [2] <strong>and</strong> the value y 2 of a variable y at the1028


27.7 HIGHER-ORDER EQUATIONSy1.00.80.6y0.40.20.20.40.60.8c−1.0−0.8−0.6−0.4−0.2−0.11.0 xFigure 27.6 The isocline method. The cross lines on each isocline show theslopes that solutions of dy/dx = −2xy must have at the points where theycross the isoclines. The heavy line is the solution with y(0) = 1, namelyexp(−x 2 ).second calculational point x 2 . The integration of these equations by the methodsdiscussed in the previous section presents no particular difficulty, provided thatall the equations are advanced through each particular step be<strong>for</strong>e any of themis taken through the following step.Higher-order equations in one dependent <strong>and</strong> one independent variable can bereduced to a set of simultaneous equations, provided that they can be written inthe <strong>for</strong>md R ydx R = f(x, y, y′ ,...,y (R−1) ), (27.82)where R is the order of the equation. To do this, a new set of variables p [r] isdefined byp [r] = dr y, r =1, 2,...,R− 1. (27.83)dxr Equation (27.82) is then equivalent to the following set of simultaneous first-orderequations:dydx = p [1],dp [r]dx = p [r+1], r =1, 2,...,R− 2, (27.84)dp [R−1]= f(x, y, p [1] ,...,p [R−1] ).dx1029


NUMERICAL METHODSThese can then be treated in the way indicated in the previous paragraph. Theextension to more than one dependent variable is straight<strong>for</strong>ward.In practical problems it often happens that boundary conditions applicableto a higher-order equation consist not of the values of the function <strong>and</strong> all itsderivatives at one particular point but of, say, the values of the function at twoseparate end-points. In these cases a solution cannot be found using an explicitstep-by-step ‘marching’ scheme, in which the solutions at successive values of theindependent variable are calculated using solution values previously found. Othermethods have to be tried.One obvious method is to treat the problem as a ‘marching one’, but to use anumber of (intelligently guessed) initial values <strong>for</strong> the derivatives at the startingpoint. The aim is then to find, by interpolation or some other <strong>for</strong>m of iteration,those starting values <strong>for</strong> the derivatives that will produce the given value of thefunction at the finishing point.In some cases the problem can be reduced by a differencing scheme to a matrixequation. Such a case is that of a second-order equation <strong>for</strong> y(x) with constantcoefficients <strong>and</strong> given values of y at the two end-points. Consider the second-orderequationwith the boundary conditionsy ′′ +2ky ′ + µy = f(x), (27.85)y(0) = A, y(1) = B.If (27.85) is replaced by a central difference equation,y i+1 − 2y i + y i−1h 2we obtain from it the recurrence relation+2k y i+1 − y i−12h+ µy i = f(x i ),(1 + kh)y i+1 +(µh 2 − 2)y i +(1− kh)y i−1 = h 2 f(x i ).For h =1/(N − 1) this is in exactly the <strong>for</strong>m of the N × N tridiagonal matrixequation (27.30), withb 1 = b N =1, c 1 = a N =0,a i =1− kh, b i = µh 2 − 2, c i =1+kh, i =2, 3,...,N− 1,<strong>and</strong> y 1 replaced by A, y N by B <strong>and</strong> y i by h 2 f(x i )<strong>for</strong>i =2, 3,...,N − 1. Thesolutions can be obtained as in (27.31) <strong>and</strong> (27.32).27.8 Partial differential equationsThe extension of previous methods to partial differential equations, thus involvingtwo or more independent variables, proceeds in a more or less obvious way. Rather1030


27.8 PARTIAL DIFFERENTIAL EQUATIONSthan an interval divided into equal steps by the points at which solutions to theequations are to be found, a mesh of points in two or more dimensions has to beset up <strong>and</strong> all the variables given an increased number of subscripts.Considerations of the stability, accuracy <strong>and</strong> feasibility of particular calculationalschemes are the same as <strong>for</strong> the one-dimensional case in principle, but inpractice are too complicated to be discussed here.Rather than note generalities that we are unable to pursue in any quantitativeway, we will conclude this chapter by indicating in outline how two familiarpartial differential equations of physical science can be set up <strong>for</strong> numericalsolution. The first of these is Laplace’s equation in two dimensions,∂ 2 φ∂x 2 + ∂2 φ=0, (27.86)∂y2 the value of φ being given on the perimeter of a closed domain.A grid with spacings ∆x <strong>and</strong> ∆y in the two directions is first chosen, so that,<strong>for</strong> example, x i st<strong>and</strong>s <strong>for</strong> the point x 0 + i∆x <strong>and</strong> φ i,j <strong>for</strong> the value φ(x i ,y j ). Next,using a second central difference <strong>for</strong>mula, (27.86) is turned intoφ i+1,j − 2φ i,j + φ i−1,j(∆x) 2 + φ i,j+1 − 2φ i,j + φ i,j−1(∆y) 2 =0,(27.87)<strong>for</strong> i =0, 1,...,N <strong>and</strong> j =0, 1,...,M.If(∆x) 2 = λ(∆y) 2 then this becomes therecurrence relationshipφ i+1,j + φ i−1,j + λ(φ i,j+1 + φ i,j−1 )=2(1+λ)φ i,j . (27.88)The boundary conditions in their simplest <strong>for</strong>m (i.e. <strong>for</strong> a rectangular domain)mean thatφ 0,j , φ N,j , φ i,0 , φ i,M (27.89)have predetermined values. Non-rectangular boundaries can be accommodated,either by more complex boundary-value prescriptions or by using non-Cartesiancoordinates.To find a set of values satisfying (27.88), an initial guess of a complete set ofvalues <strong>for</strong> the φ i,j is made, subject to the requirement that the quantities listedin (27.89) have the given fixed values; those values that are not on the boundaryare then adjusted iteratively in order to try to bring about condition (27.88)everywhere. Clearly one scheme is to set λ = 1 <strong>and</strong> recalculate each φ i,j as themean of the four current values at neighbouring grid-points, using (27.88) directly,<strong>and</strong> then to iterate this recalculation until no value of φ changes significantlyafter a complete cycle through all values of i <strong>and</strong> j. This procedure is the simplestof such ‘relaxation’ methods; <strong>for</strong> a slightly more sophisticated scheme see exercise27.26 at the end of this chapter. The reader is referred to specialist books <strong>for</strong>fuller accounts of how this approach can be made faster <strong>and</strong> more accurate.1031


NUMERICAL METHODSOur final example is based upon the one-dimensional diffusion equation <strong>for</strong>the temperature φ of a system:∂φ∂t = φκ∂2 ∂x 2 . (27.90)If φ i,j st<strong>and</strong>s <strong>for</strong> φ(x 0 + i∆x, t 0 + j∆t) then a <strong>for</strong>ward difference representationof the time derivative <strong>and</strong> a central difference representation <strong>for</strong> the spatialderivative lead to the following relationship:φ i,j+1 − φ i,j∆t= κ φ i+1,j − 2φ i,j + φ i−1,j(∆x) 2 . (27.91)This allows the construction of an explicit scheme <strong>for</strong> generating the temperaturedistribution at later times, given that it is known at some earlier time:φ i,j+1 = α(φ i+1,j + φ i−1,j )+(1− 2α)φ i,j , (27.92)where α = κ∆t/(∆x) 2 .Although this scheme is explicit, it is not a good one because of the asymmetricway in which the differences are <strong>for</strong>med. However, the effect of this can beminimised if we study <strong>and</strong> correct <strong>for</strong> the errors introduced in the following way.Taylor’s series <strong>for</strong> the time variable givesφ i,j+1 = φ i,j +∆t ∂φ i,j+ (∆t)2 ∂ 2 φ i,j∂t 2! ∂t 2 + ··· , (27.93)using the same notation as previously. Thus the first correction term to the LHSof (27.91) is− ∆t ∂ 2 φ i,j2 ∂t 2 . (27.94)The first term omitted on the RHS of the same equation is, by a similar argument,−κ 2(∆x)2 ∂ 4 φ i,j4! ∂x 4 . (27.95)But, using the fact that φ satisfies (27.90), we obtain∂ 2 φ∂t 2= ∂ ∂t( ) ( )κ ∂2 φ∂x 2 = κ ∂2 ∂φ∂x 2 = κ 2 ∂4 φ∂t ∂x 4 , (27.96)<strong>and</strong> so, to this accuracy, the two errors (27.94) <strong>and</strong> (27.95) can be made to cancelif α is chosen such that− κ2 ∆t2= − 2κ(∆x)2 , i.e. α = 1 4!6 .1032


27.9 EXERCISES27.9 Exercises27.1 Use an iteration procedure to find the root of the equation 40x =expx to foursignificant figures.27.2 Using the Newton–Raphson procedure find, correct to three decimal places, theroot nearest to 7 of the equation 4x 3 +2x 2 − 200x − 50 = 0.27.3 Show the following results about rearrangement schemes <strong>for</strong> polynomial equations.(a) That if a polynomial equation g(x) ≡ x m − f(x) = 0, where f(x) is apolynomial of degree less than m <strong>and</strong> <strong>for</strong> which f(0) ≠ 0, is solved usinga rearrangement iteration scheme x n+1 =[f(x n )] 1/m , then, in general, thescheme will have only first-order convergence.(b) By considering the cubic equationx 3 − ax 2 +2abx − (b 3 + ab 2 )=0<strong>for</strong> arbitrary non-zero values of a <strong>and</strong> b, demonstrate that, in special cases, thesame rearrangement scheme can give second- (or higher-) order convergence.27.4 The square root of a number N is to be determined by means of the iterationscheme[ ( ) ]x n+1 = x n 1 − N − x2n f(N) .Determine how to choose f(N) so that the process has second-order convergence.Given that √ 7 ≈ 2.65, calculate √ 7 as accurately as a single application of the<strong>for</strong>mula will allow.27.5 Solve the following set of simultaneous equations using Gaussian elimination(including interchange where it is <strong>for</strong>mally desirable):x 1 +3x 2 +4x 3 +2x 4 =0,2x 1 +10x 2 − 5x 3 + x 4 =6,4x 2 +3x 3 +3x 4 =20,−3x 1 +6x 2 +12x 3 − 4x 4 =16.27.6 The following table of values of a polynomial p(x) of low degree contains anerror. Identify <strong>and</strong> correct the erroneous value <strong>and</strong> extend the table up to x =1.2.x p(x) x p(x)0.0 0.000 0.5 0.1650.1 0.011 0.6 0.2160.2 0.040 0.7 0.2450.3 0.081 0.8 0.2560.4 0.128 0.9 0.24327.7 Simultaneous linear equations that result in tridiagonal matrices can sometimesbe treated as three-term recurrence relations, <strong>and</strong> their solution may be foundin a similar manner to that described in chapter 15. Consider the tridiagonalsimultaneous equationsx i−1 +4x i + x i+1 =3(δ i+1,0 − δ i−1,0 ), i =0, ±1, ±2,... .Prove that, <strong>for</strong> i>0, the equations have a general solution of the <strong>for</strong>m x i =αp i + βq i ,wherep <strong>and</strong> q are the roots of a certain quadratic equation. Show thata similar result holds <strong>for</strong> i


NUMERICAL METHODS27.8 A possible rule <strong>for</strong> obtaining an approximation to an integral is the mid-pointrule, givenby∫ x0 +∆xf(x) dx =∆xf(x 0 + 1 2 ∆x)+O(∆x3 ).x 0Writing h <strong>for</strong> ∆x, <strong>and</strong> evaluating all derivates at the mid-point of the interval(x, x +∆x), use a Taylor series expansion to find, up to O(h 5 ), the coefficients ofthe higher-order errors in both the trapezium <strong>and</strong> mid-point rules. Hence find alinear combination of these two rules that gives O(h 5 ) accuracy <strong>for</strong> each step ∆x.27.9 Although it can easily be shown, by direct calculation, that∫ ∞e −x cos(kx) dx = 101+k , 2the <strong>for</strong>m of the integr<strong>and</strong> is appropriate <strong>for</strong> Gauss–Laguerre numerical integration.Using a 5-point <strong>for</strong>mula, investigate the range of values of k <strong>for</strong> which the<strong>for</strong>mula gives accurate results. At about what value of k do the results becomeinaccurate at the 1% level?27.10 Using the points <strong>and</strong> weights given in table 27.9, answer the following questions.(a) A table of unnormalised Hermite polynomials H n (x) has been spattered withink blots <strong>and</strong> gives H 5 (x) as32x 5 −?x 3 + 120x <strong>and</strong> H 4 (x) as?x 4 −?x 2 + 12,where the coefficients marked ? cannot be read. What should they read?(b) What is the value of the integral∫ ∞e −2x2I =−∞ 4x 2 +3x +1 dx,as given by a 7-point integration routine?27.11 Consider the integrals I p defined byI p =∫ 1−1x 2p√1 − x2 dx.(a) By setting x =sinθ <strong>and</strong> using the results given in exercise 2.42, show that I phas the valueI p =2 2p − 1 2p − 32p 2p − 2 ··· 1 π2 2 .(b) Evaluate I p <strong>for</strong> p =1, 2,... ,6 using 5- <strong>and</strong> 6-point Gauss–Chebyshev integration(convenientlyrunonaspreadsheetsuchasExcel)<strong>and</strong> compare theresults with those in (a). In particular, show that, as expected, the 5-pointscheme first fails to be accurate when the order of the polynomial numerator(2p) exceeds (2×5)−1 = 9. Likewise, verify that the 6-point scheme evaluatesI 5 accurately but is in error <strong>for</strong> I 6 .27.12 In normal use only a single application of n-point Gaussian quadrature is made,using a value of n that is estimated from experience to be ‘safe’. However, it isinstructive to examine what happens when n is changed in a controlled way.(a) Evaluate the integral∫ 5 √I n = 7x − x2 − 10 dx2using n-point Gauss–Legendre <strong>for</strong>mulae <strong>for</strong> n =2, 3,... ,6. Estimate (to 4s.f.) the value I ∞ you would obtain <strong>for</strong> very large n <strong>and</strong> compare it with theresult I obtained by exact integration. Explain why the variation of I n withn is monotonically decreasing.1034


27.9 EXERCISES(b) Try to repeat the processes described in (a) <strong>for</strong> the integrals∫ 51J n = √7x − x2 − 10 dx.2WhyisitverydifficulttoestimateJ ∞ ?27.13 Given a r<strong>and</strong>om number η uni<strong>for</strong>mly distributed on (0, 1), determine the functionξ = ξ(η) that would generate a r<strong>and</strong>om number ξ distributed as(a) 2ξ√ on 0 ≤ ξ


NUMERICAL METHODS(b) Substitute them into the predictor equation <strong>and</strong>, by making that expression<strong>for</strong> ȳ n+1 coincide with the true Taylor series <strong>for</strong> y n+1 up to order h 3 , establishsimultaneous equations that determine the values of a 1 ,a 2 <strong>and</strong> a 3 .(c) Find the Taylor series <strong>for</strong> f n+1 <strong>and</strong> substitute it <strong>and</strong> that <strong>for</strong> f n−1 into thecorrector equation. Make the corrected prediction <strong>for</strong> y n+1 coincide with thetrue Taylor series by choosing the weights b 1 ,b 2 <strong>and</strong> b 3 appropriately.(d) The values of the numerical solution of the differential equationdy 2(1 + x)y + x3/2=dx 2x(1 + x)at three values of x are given in the following table:x 0.1 0.2 0.3y(x) 0.030 628 0.084 107 0.150 328Use the above predictor–corrector scheme to find the value of y(0.4) <strong>and</strong>compare your answer with the accurate value, 0.225 577.27.18 If dy/dx = f(x, y) then show thatd 2 fdx = ∂2 f2 ∂x +2f ∂2 f2 ∂x∂y + ∂2 ff2∂y + ∂f ( ) 2∂f ∂f2 ∂x ∂y + f .∂yHence verify, by substitution <strong>and</strong> the subsequent expansion of arguments inTaylor series of their own, that the scheme given in (27.79) coincides with theTaylor expansion (27.68), i.e.y i+1 = y i + hy (1)i+ h2iup to terms in h 3 .27.19 To solve the ordinary differential equation2! y(2)+ h33! y(3) i+ ··· ,du= f(u, t)dt<strong>for</strong> f = f(t), the explicit two-step finite difference schemeu n+1 = αu n + βu n−1 + h(µf n + νf n−1 )may be used. Here, in the usual notation, h is the time step, t n = nh, u n = u(t n )<strong>and</strong> f n = f(u n ,t n ); α, β, µ, <strong>and</strong>ν are constants.(a) A particular scheme has α =1,β =0,µ=3/2 <strong>and</strong>ν = −1/2. By consideringTaylor expansions about t = t n <strong>for</strong> both u n+j <strong>and</strong> f n+j , show that this schemegives errors of order h 3 .(b) Find the values of α, β, µ <strong>and</strong> ν that will give the greatest accuracy.27.20 Set up a finite difference scheme to solve the ordinary differential equationx d2 φdx + dφ2 dx =0in the range 1 ≤ x ≤ 4, subject to the boundary conditions φ(1) = 2 <strong>and</strong>dφ/dx =2atx =4.UsingN equal increments, ∆x, inx, obtain the generaldifference equation <strong>and</strong> state how the boundary conditions are incorporatedinto the scheme. Setting ∆x equal to the (crude) value 1, obtain the relevantsimultaneous equations <strong>and</strong> so obtain rough estimates <strong>for</strong> φ(2),φ(3) <strong>and</strong> φ(4).Finally, solve the original equation analytically <strong>and</strong> compare your numericalestimates with the accurate values.1036


27.9 EXERCISES27.21 Write a computer program that would solve, <strong>for</strong> a range of values of λ, thedifferential equationdydx = 1√ , y(0) = 1,x2 + λy2 using a third-order Runge–Kutta scheme. Consider the difficulties that mightarise when λ


NUMERICAL METHODS27.25 Laplace’s equation,∂ 2 V∂x + ∂2 V2 ∂y =0, 2is to be solved <strong>for</strong> the region <strong>and</strong> boundary conditions shown in figure 27.7.V =80−∞40404040404040∞202020V =0Figure 27.7Region, boundary values <strong>and</strong> initial guessed solution values.Starting from the given initial guess <strong>for</strong> the potential values V , <strong>and</strong> using thesimplest possible <strong>for</strong>m of relaxation, obtain a better approximation to the actualsolution. Do not aim to be more accurate than ± 0.5 units, <strong>and</strong> so terminate theprocess when subsequent changes would be no greater than this.27.26 Consider the solution, φ(x, y), of Laplace’s equation in two dimensions using arelaxation method on a square grid with common spacing h. Asinthemaintext,denote φ(x 0 + ih, y 0 + jh) byφ i,j . Further, define φ m,ni,jbyφ m,ni,j≡ ∂m+n φ∂x m ∂y nevaluated at (x 0 + ih, y 0 + jh).(a) Show thatφ 4,0i,j+2φ 2,2i,j+ φ 0,4i,j=0.(b) Working up to terms of order h 5 , find Taylor series expansions, expressed interms of the φ m,ni,j ,<strong>for</strong> S ±,0 = φ i+1,j + φ i−1,j ,S 0,± = φ i,j+1 + φ i,j−1 .(c) Find a corresponding expansion, to the same order of accuracy, <strong>for</strong> φ i±1,j+1 +φ i±1,j−1 <strong>and</strong> hence show thatS ±,± = φ i+1,j+1 + φ i+1,j−1 + φ i−1,j+1 + φ i−1,j−1has the <strong>for</strong>m)+h44φ 0,0i,j+2h 2 (φ 2,0i,j+ φ 0,2i,j6 (φ4,0 i,j+6φ 2,2i,j+ φ 0,4i,j ).(d) Evaluate the expression 4(S ±,0 +S 0,± )+S ±,± <strong>and</strong> hence deduce that a possiblerelaxation scheme, good to the fifth order in h, is to recalculate each φ i,j asthe weighted mean of the current values of its four nearest neighbours (eachwith weight 1 1) <strong>and</strong> its four next-nearest neighbours (each with weight ).5 201038


27.10 HINTS AND ANSWERS27.27 The Schrödinger equation <strong>for</strong> a quantum mechanical particle of mass m movingin a one-dimensional harmonic oscillator potential V (x) =kx 2 /2is− 2 d 2 ψ2m dx + kx2 ψ= Eψ.2 2For physically acceptable solutions, the wavefunction ψ(x) must be finite at x =0,tend to zero as x →±∞<strong>and</strong>benormalised,sothat ∫ |ψ| 2 dx =1.Inpractice,these constraints mean that only certain (quantised) values of E, the energy ofthe particle, are allowed. The allowed values fall into two groups: those <strong>for</strong> whichψ(0) = 0 <strong>and</strong> those <strong>for</strong> which ψ(0) ≠0.Show that if the unit of length is taken as [ 2 /(mk)] 1/4 <strong>and</strong> the unit of energyis taken as (k/m) 1/2 , then the Schrödinger equation takes the <strong>for</strong>md 2 ψdy 2 +(2E′ − y 2 )ψ =0.Devise an outline computerised scheme, using Runge–Kutta integration, that willenable you to:(a) determine the three lowest allowed values of E;(b) tabulate the normalised wavefunction corresponding to the lowest allowedenergy.You should consider explicitly:(i) the variables to use in the numerical integration;(ii) how starting values near y = 0 are to be chosen;(iii) how the condition on ψ as y →±∞is to be implemented;(iv) how the required values of E are to be extracted from the results of theintegration;(v) how the normalisation is to be carried out.27.10 Hints <strong>and</strong> answers27.1 5.370.27.3 (a) ξ ≠0<strong>and</strong>f ′ (ξ) ≠ 0 in general; (b) ξ = b, but f ′ (b) = 0 whilst f(b) ≠0.27.5 Interchange is <strong>for</strong>mally needed <strong>for</strong> the first two steps, though in this case no errorwill result if it is not carried out; x 1 = −12, x 2 =2,x 3 = −1, x 4 =5.27.7 The quadratic equation is z 2 +4z +1=0;α + β − 3=x 0 = α ′ + β ′ +3.With p = −2+ √ 3<strong>and</strong>q = −2 − √ 3, β must be zero <strong>for</strong> i>0<strong>and</strong>α ′ must bezero <strong>for</strong> i0, x i =0<strong>for</strong>i =0,x i = −3(−2 − √ 3) i<strong>for</strong> i0; ξ =ln2η <strong>for</strong> 0


NUMERICAL METHODSV =80−∞40.541.8 46.7 48.4 46.741.840.5∞16.8 20.4 16.8V =0Figure 27.8 The solution to exercise 27.25.27.15 1 − x 2 /2+x 4 /8 − x 6 /48; 1.0000, 0.9950, 0.9802, 0.9560, 0.9231, 0.8825; exactsolution y =exp(−x 2 /2).27.17 (b) a 1 =23/12, a 2 = −4/3, a 3 =5/12.(c) b 1 =5/12, b 2 =2/3, b 3 = −1/12.(d) ȳ(0.4) = 0.224 582, y(0.4) = 0.225 527 after correction.27.19 (a) The error is 5h 3 u (3)n /12 + O(h 4 ).(b) α = −4, β =5,µ=4<strong>and</strong>ν =227.21 For λ positive the solutions are (boringly) monotonic functions of x. Withy(0)given, there are no real solutions at all <strong>for</strong> any negative λ!27.23 (a) Setting A∆t = c∆x gives, <strong>for</strong> example, u(0, 2) = (1 − c) 2 , u(1, 2) = 2c(1 − c),u(2, 2) = c 2 . For stability, 0


28Group theoryFor systems that have some degree of symmetry, full exploitation of that symmetryis desirable. Significant physical results can sometimes be deduced simply by astudy of the symmetry properties of the system under investigation. Consequentlyit becomes important, <strong>for</strong> such a system, to identify all those operations (rotations,reflections, inversions) that carry the system into a physically indistinguishablecopy of itself.The study of the properties of the complete set of such operations <strong>for</strong>msone application of group theory. Though this is the aspect of most interest tothe physical scientist, group theory itself is a much larger subject <strong>and</strong> of greatimportance in its own right. Consequently we leave until the next chapter anydirect applications of group theoretical results <strong>and</strong> concentrate on building upthe general mathematical properties of groups.28.1 GroupsAs an example of symmetry properties, let us consider the sets of operations,such as rotations, reflections, <strong>and</strong> inversions, that trans<strong>for</strong>m physical objects, <strong>for</strong>example molecules, into physically indistinguishable copies of themselves, so thatonly the labelling of identical components of the system (the atoms) changes inthe process. For differently shaped molecules there are different sets of operations,but in each case it is a well-defined set, <strong>and</strong> with a little practice all members ofeach set can be identified.As simple examples, consider (a) the hydrogen molecule, <strong>and</strong> (b) the ammoniamolecule illustrated in figure 28.1. The hydrogen molecule consists of two atomsH of hydrogen <strong>and</strong> is carried into itself by any of the following operations:(i) any rotation about its long axis;(ii) rotation through π about an axis perpendicular to the long axis <strong>and</strong>passing through the point M that lies midway between the atoms;1041


GROUP THEORYNHMHHH(a)(b)HFigure 28.1(a) The hydrogen molecule, <strong>and</strong> (b) the ammonia molecule.(iii) inversion through the point M;(iv) reflection in the plane that passes through M <strong>and</strong> has its normal parallelto the long axis.These operations collectively <strong>for</strong>m the set of symmetry operations <strong>for</strong> the hydrogenmolecule.The somewhat more complex ammonia molecule consists of a tetrahedron withan equilateral triangular base at the three corners of which lie hydrogen atomsH, whilst a nitrogen atom N is sited at the fourth vertex of the tetrahedron. Theset of symmetry operations on this molecule is limited to rotations of π/3 <strong>and</strong>2π/3 about the axis joining the centroid of the equilateral triangle to the nitrogenatom, <strong>and</strong> reflections in the three planes containing that axis <strong>and</strong> each of thehydrogen atoms in turn. However, if the nitrogen atom could be replaced by afourth hydrogen atom, <strong>and</strong> all interatomic distances equalised in the process, thenumber of symmetry operations would be greatly increased.Once all the possible operations in any particular set have been identified, itmust follow that the result of applying two such operations in succession will beidentical to that obtained by the sole application of some third (usually different)operation in the set – <strong>for</strong> if it were not, a new member of the set would havebeen found, contradicting the assumption that all members have been identified.Such observations introduce two of the main considerations relevant to decidingwhether a set of objects, here the rotation, reflection <strong>and</strong> inversion operations,qualifies as a group in the mathematically tightly defined sense. These two considerationsare (i) whether there is some law <strong>for</strong> combining two members of the set,<strong>and</strong> (ii) whether the result of the combination is also a member of the set. Theobvious rule of combination has to be that the second operation is carried outon the system that results from application of the first operation, <strong>and</strong> we havealready seen that the second requirement is satisfied by the inclusion of all suchoperations in the set. However, <strong>for</strong> a set to qualify as a group, more than thesetwo conditions have to be satisfied, as will now be made clear.1042


28.1 GROUPS28.1.1 Definition of a groupAgroupG is a set of elements {X,Y,...}, together with a rule <strong>for</strong> combiningthem that associates with each ordered pair X, Y a ‘product’ or combination lawX • Y <strong>for</strong> which the following conditions must be satisfied.(i) For every pair of elements X, Y that belongs to G, the product X • Y alsobelongs to G. (This is known as the closure property of the group.)(ii) For all triples X, Y , Z the associative law holds; in symbols,X • (Y • Z) =(X • Y ) • Z. (28.1)(iii) There exists a unique element I, belonging to G, with the property thatI • X = X = X • I (28.2)<strong>for</strong> all X belonging to G. This element I is known as the identity elementof the group.(iv) For every element X of G, there exists an element X −1 , also belonging toG, such thatX −1 is called the inverse of X.X −1 • X = I = X • X −1 . (28.3)An alternative notation in common use is to write the elements of a group Gas the set {G 1 ,G 2 ,...} or, more briefly, as {G i }, a typical element being denotedby G i .It should be noticed that, as given, the nature of the operation • is not stated. Itshould also be noticed that the more general term element, rather than operation,has been used in this definition. We will see that the general definition of agroup allows as elements not only sets of operations on an object but also sets ofnumbers, of functions <strong>and</strong> of other objects, provided that the interpretation of •is appropriately defined.In one of the simplest examples of a group, namely the group of all integersunder addition, the operation • is taken to be ordinary addition. In this group therole of the identity I is played by the integer 0, <strong>and</strong> the inverse of an integer X is−X. That requirements (i) <strong>and</strong> (ii) are satisfied by the integers under addition istrivially obvious. A second simple group, under ordinary multiplication, is <strong>for</strong>medby the two numbers 1 <strong>and</strong> −1; in this group, closure is obvious, 1 is the identityelement, <strong>and</strong> each element is its own inverse.It will be apparent from these two examples that the number of elements in agroup can be either finite or infinite. In the <strong>for</strong>mer case the group is called a finitegroup <strong>and</strong> the number of elements it contains is called the order of the group,which we will denote by g; an alternative notation is |G| but has obvious dangers1043


GROUP THEORYif matrices are involved. In the notation in which G = {G 1 ,G 2 ,...,G n } the orderof the group is clearly n.As we have noted, <strong>for</strong> the integers under addition zero is the identity. Forthe group of rotations <strong>and</strong> reflections, the operation of doing nothing, i.e. thenull operation, plays this role. This latter identification may seem artificial, butit is an operation, albeit trivial, which does leave the system in a physicallyindistinguishable state, <strong>and</strong> needs to be included. One might add that without itthe set of operations would not <strong>for</strong>m a group <strong>and</strong> none of the powerful resultswe will derive later in this <strong>and</strong> the next chapter could be justifiably applied togive deductions of physical significance.In the examples of rotations <strong>and</strong> reflections mentioned earlier, • has been takento mean that the left-h<strong>and</strong> operation is carried out on the system that resultsfrom application of the right-h<strong>and</strong> operation. ThusZ = X • Y (28.4)means that the effect on the system of carrying out Z isthesameaswouldbe obtained by first carrying out Y <strong>and</strong> then carrying out X. The order of theoperations should be noted; it is arbitrary in the first instance but, once chosen,must be adhered to. The choice we have made is dictated by the fact that mostof our applications involve the effect of rotations <strong>and</strong> reflections on functions ofspace coordinates, <strong>and</strong> it is usual, <strong>and</strong> our practice in the rest of this book, towrite operators acting on functions to the left of the functions.It will be apparent that <strong>for</strong> the above-mentioned group, integers under ordinaryaddition, it is true thatY • X = X • Y (28.5)<strong>for</strong> all pairs of integers X, Y . If any two particular elements of a group satisfy(28.5), they are said to commute under the operation •; if all pairs of elements ina group satisfy (28.5), then the group is said to be Abelian. Thesetofallintegers<strong>for</strong>ms an infinite Abelian group under (ordinary) addition.As we show below, requirements (iii) <strong>and</strong> (iv) of the definition of a groupare over-dem<strong>and</strong>ing (but self-consistent), since in each of equations (28.2) <strong>and</strong>(28.3) the second equality can be deduced from the first by using the associativityrequired by (28.1). The mathematical steps in the following arguments are allvery simple, but care has to be taken to make sure that nothing that has notyet been proved is used to justify a step. For this reason, <strong>and</strong> to act as a modelin logical deduction, a reference in Roman numerals to the previous result,or to the group definition used, is given over each equality sign. Such explicitdetailed referencing soon becomes tiresome, but it should always be available ifneeded.1044


28.1 GROUPS◮Using only the first equalities in (28.2) <strong>and</strong> (28.3), deduce the second ones.Consider the expression X −1 • (X • X −1 );X −1 • (X • X −1 ) (ii) =(X −1 −1 (iv)• X) • X = I • X −1(iii)= X −1 . (28.6)But X −1 belongs to G, <strong>and</strong> so from (iv) there is an element U in G such thatU • X −1 = I.(v)Form the product of U with the first <strong>and</strong> last expressions in (28.6) to giveU • (X −1 • (X • X −1 −1 (v))) = U • X = I. (28.7)Trans<strong>for</strong>ming the left-h<strong>and</strong> side of this equation givesU • (X −1 • (X • X −1 )) (ii) =(U • X −1 ) • (X • X −1 )(v)= I • (X • X −1 )(iii)= X • X −1 . (28.8)Comparing (28.7), (28.8) shows thatX • X −1 = I, (iv) ′i.e. the second equality in group definition (iv). SimilarlyX • I (iv) = X • (X −1 • X) (ii) =(X • X −1 ) • X(iv) ′= I • X(iii)= X. (iii ′ )i.e. the second equality in group definition (iii). ◭The uniqueness of the identity element I canalsobedemonstratedratherthanassumed. Suppose that I ′ , belonging to G, also has the propertyTake X as I, thenFurther, from (iii ′ ),I ′ • X = X = X • I ′ <strong>for</strong> all X belonging to G.I ′ • I = I. (28.9)X = X • I <strong>for</strong> all X belonging to G,1045


GROUP THEORY<strong>and</strong> setting X = I ′ givesI ′ = I ′ • I. (28.10)It then follows from (28.9), (28.10) that I = I ′ , showing that in any particulargroup the identity element is unique.In a similar way it can be shown that the inverse of any particular elementis unique. If U <strong>and</strong> V are two postulated inverses of an element X of G, byconsidering the productU • (X • V )=(U • X) • V,it can be shown that U = V . The proof is left to the reader.Given the uniqueness of the inverse of any particular group element, it followsthat(U • V • ···• Y • Z) • (Z −1 • Y −1 • ···• V −1 • U −1 )=(U • V • ···• Y ) • (Z • Z −1 ) • (Y −1 • ···• V −1 • U −1 )=(U • V • ···• Y ) • (Y −1 • ···• V −1 • U −1 )..= I,where use has been made of the associativity <strong>and</strong> of the two equations Z • Z −1 = I<strong>and</strong> I • X = X. Thus the inverse of a product is the product of the inverses inreverse order, i.e.(U • V • ···• Y • Z) −1 =(Z −1 • Y −1 • ···• V −1 • U −1 ). (28.11)Further elementary results that can be obtained by arguments similar to thoseabove are as follows.(i) Given any pair of elements X, Y belonging to G, there exist uniqueelements U, V, also belonging to G, such thatX • U = Y <strong>and</strong> V • X = Y.Clearly U = X −1 • Y ,<strong>and</strong>V = Y • X −1 , <strong>and</strong> they can be shown to beunique. This result is sometimes called the division axiom.(ii) The cancellation law can be stated as follows. IfX • Y = X • Z<strong>for</strong> some X belonging to G, thenY = Z. Similarly,implies the same conclusion.Y • X = Z • X1046


28.1 GROUPSLMKFigure 28.2 Reflections in the three perpendicular bisectors of the sides ofan equilateral triangle take the triangle into itself.(iii) Forming the product of each element of G with a fixed element X of Gsimply permutes the elements of G; this is often written symbolically asG • X = G. If this were not so, <strong>and</strong> X • Y <strong>and</strong> X • Z were not differenteven though Y <strong>and</strong> Z were, application of the cancellation law would leadto a contradiction. This result is called the permutation law.In any finite group of order g, any element X when combined with itself to<strong>for</strong>m successively X 2 = X • X, X 3 = X • X 2 , ... will, after at most g − 1suchcombinations, produce the group identity I. OfcourseX 2 , X 3 , ... are some ofthe original elements of the group, <strong>and</strong> not new ones. If the actual number ofcombinations needed is m−1, i.e. X m = I, thenm is called the order of the elementX in G. The order of the identity of a group is always 1, <strong>and</strong> that of any otherelement of a group that is its own inverse is always 2.◮Determine the order of the group of (two-dimensional) rotations <strong>and</strong> reflections that takea plane equilateral triangle into itself <strong>and</strong> the order of each of the elements. The group isusually known as 3m (to physicists <strong>and</strong> crystallographers) or C 3v (to chemists).There are two (clockwise) rotations, by 2π/3 <strong>and</strong>4π/3, about an axis perpendicular tothe plane of the triangle. In addition, reflections in the perpendicular bisectors of the threesides (see figure 28.2) have the defining property. To these must be added the identityoperation. Thus in total there are six distinct operations <strong>and</strong> so g = 6 <strong>for</strong> this group.To reproduce the identity operation either of the rotations has to be applied three times,whilst any of the reflections has to be applied just twice in order to recover the originalsituation. Thus each rotation element of the group has order 3, <strong>and</strong> each reflection elementhas order 2. ◭A so-called cyclic group is one <strong>for</strong> which all members of the group can begenerated from just one element X (say). Thus a cyclic group of order g can bewritten asG = { I,X,X 2 ,X 3 ,...,X g−1} .1047


GROUP THEORYIt is clear that cyclic groups are always Abelian <strong>and</strong> that each element, apartfrom the identity, has order g, the order of the group itself.28.1.2 Further examples of groupsIn this section we consider some sets of objects, each set together with a law ofcombination, <strong>and</strong> investigate whether they qualify as groups <strong>and</strong>, if not, why not.We have already seen that the integers <strong>for</strong>m a group under ordinary addition,but it is immediately apparent that (even if zero is excluded) they do not doso under ordinary multiplication. Unity must be the identity of the set, but therequisite inverse of any integer n, namely1/n, does not belong to the set ofintegers <strong>for</strong> any n other than unity.Other infinite sets of quantities that do <strong>for</strong>m groups are the sets of all realnumbers, or of all complex numbers, under addition, <strong>and</strong> of the same two setsexcluding 0 under multiplication. All these groups are Abelian.Although subtraction <strong>and</strong> division are normally considered the obvious counterpartsof the operations of (ordinary) addition <strong>and</strong> multiplication, they are notacceptable operations <strong>for</strong> use within groups since the associative law, (28.1), doesnot hold. Explicitly,X − (Y − Z) ≠(X − Y ) − Z,X ÷ (Y ÷ Z) ≠(X ÷ Y ) ÷ Z.From within the field of all non-zero complex numbers we can select just thosethat have unit modulus, i.e. are of the <strong>for</strong>m e iθ where 0 ≤ θ


28.2 FINITE GROUPS28.2 Finite groupsWhilst many properties of physical systems (e.g. angular momentum) are relatedto the properties of infinite, <strong>and</strong>, in particular, continuous groups, the symmetryproperties of crystals <strong>and</strong> molecules are more intimately connected with those offinite groups. We there<strong>for</strong>e concentrate in this section on finite sets of objects thatcan be combined in a way satisfying the group postulates.Although it is clear that the set of all integers does not <strong>for</strong>m a group underordinary multiplication, restricted sets can do so if the operation involved is multiplication(mod N) <strong>for</strong> suitable values of N; this operation will be explained below.As a simple example of a group with only four members, consider the set Sdefined as follows:S = {1, 3, 5, 7} under multiplication (mod 8).To find the product (mod 8) of any two elements, we multiply them together inthe ordinary way, <strong>and</strong> then divide the answer by 8, treating the remainder afterdoing so as the product of the two elements. For example, 5 × 7 = 35, which ondividing by 8 gives a remainder of 3. Clearly, since Y × Z = Z × Y , the full setof different products is1 × 1=1, 1 × 3=3, 1 × 5=5, 1 × 7=7,3 × 3=1, 3 × 5=7, 3 × 7=5,5 × 5=1, 5 × 7=3,7 × 7=1.The first thing to notice is that each multiplication produces a member of theoriginal set, i.e. the set is closed. Obviously the element 1 takes the role of theidentity, i.e. 1 × Y = Y <strong>for</strong> all members Y of the set. Further, <strong>for</strong> each element Yof the set there is an element Z (equal to Y , as it happens, in this case) such thatY × Z = 1, i.e. each element has an inverse. These observations, together with theassociativity of multiplication (mod 8), show that the set S is an Abelian groupof order 4.It is convenient to present the results of combining any two elements of agroup in the <strong>for</strong>m of multiplication tables – akin to those which used to appear inelementary arithmetic books be<strong>for</strong>e electronic calculators were invented! Writtenin this much more compact <strong>for</strong>m the above example is expressed by table 28.1.Although the order of the two elements being combined does not matter herebecause the group is Abelian, we adopt the convention that if the product in ageneral multiplication table is written X • Y then X is taken from the left-h<strong>and</strong>column <strong>and</strong> Y is taken from the top row. Thus the bold ‘7’ in the table is theresult of 3 × 5, rather than of 5 × 3.Whilst it would make no difference to the basic in<strong>for</strong>mation content in a tableto present the rows <strong>and</strong> columns with their headings in r<strong>and</strong>om orders, it is1049


GROUP THEORY1 3 5 71 1 3 5 73 3 1 7 55 5 7 1 37 7 5 3 1Table 28.1 The table of products <strong>for</strong> the elements of the group S = {1, 3, 5, 7}under multiplication (mod 8).usual to list the elements in the same order in both the vertical <strong>and</strong> horizontalheadings in any one table. The actual order of the elements in the common list,whilst arbitrary, is normally chosen to make the table have as much symmetry aspossible. This is initially a matter of convenience, but, as we shall see later, someof the more subtle properties of groups are revealed by putting next to each otherelements of the group that are alike in certain ways.Some simple general properties of group multiplication tables can be deducedimmediately from the fact that each row or column constitutes the elements ofthe group.(i) Each element appears once <strong>and</strong> only once in each row or column of thetable; this must be so since G • X = G (the permutation law) holds.(ii) The inverse of any element Y can be found by looking along the rowin which Y appears in the left-h<strong>and</strong> column (the Y th row), <strong>and</strong> notingthe element Z at the head of the column (the Zth column) in whichthe identity appears as the table entry. An immediate corollary is thatwhenever the identity appears on the leading diagonal, it indicates thatthe corresponding header element is of order 2 (unless it happens to bethe identity itself).(iii) For any Abelian group the multiplication table is symmetric about theleading diagonal.To get used to the ideas involved in using group multiplication tables, we nowconsider two more sets of integers under multiplication (mod N):S ′ = {1, 5, 7, 11} under multiplication (mod 24), <strong>and</strong>S ′′ = {1, 2, 3, 4} under multiplication (mod 5).These have group multiplication tables 28.2(a) <strong>and</strong> (b) respectively, as the readershould verify.If tables 28.1 <strong>and</strong> 28.2(a) <strong>for</strong> the groups S <strong>and</strong> S ′ are compared, it will be seenthat they have essentially the same structure, i.e if the elements are written as{I,A,B,C} in both cases, then the two tables are each equivalent to table 28.3.For S, I =1,A =3,B =5,C = 7 <strong>and</strong> the law of combination is multiplication(mod 8), whilst <strong>for</strong> S ′ , I =1,A =5,B =7,C = 11 <strong>and</strong> the law of combination1050


28.2 FINITE GROUPS(a)1 5 7 111 1 5 7 115 5 1 11 77 7 11 1 511 11 7 5 1(b)1 2 3 41 1 2 3 42 2 4 1 33 3 1 4 24 4 3 2 1Table 28.2 (a) The multiplication table <strong>for</strong> the group S ′ = {1, 5, 7, 11} undermultiplication (mod 24). (b) The multiplication table <strong>for</strong> the group S ′′ ={1, 2, 3, 4} under multiplication (mod 5).I A B CI I A B CA A I C BB B C I AC C B A ITable 28.3The common structure exemplified by tables 28.1 <strong>and</strong> 28.2(a).1 i −1 −i1 1 i −1 −ii i −1 −i 1−1 −1 −i 1 i−i −i 1 i −1Table 28.4 The group table <strong>for</strong> the set {1,i,−1, −i} under ordinary multiplicationof complex numbers.is multiplication (mod 24). However, the really important point is that the twogroups S <strong>and</strong> S ′ have equivalent group multiplication tables – they are said tobe isomorphic, a matter to which we will return more <strong>for</strong>mally in section 28.5.◮Determine the behaviour of the set of four elements{1,i,−1, −i}under the ordinary multiplication of complex numbers. Show that they <strong>for</strong>m a group <strong>and</strong>determine whether the group is isomorphic to either of the groups S (itself isomorphic toS ′ ) <strong>and</strong> S ′′ defined above.That the elements <strong>for</strong>m a group under the associative operation of complex multiplicationis immediate; there is an identity (1), each possible product generates a member of the set<strong>and</strong> each element has an inverse (1, −i, −1, i, respectively). The group table has the <strong>for</strong>mshown in table 28.4.We now ask whether this table can be made to look like table 28.3, which is thest<strong>and</strong>ardised <strong>for</strong>m of the tables <strong>for</strong> S <strong>and</strong> S ′ . Since the identity element of the group(1) will have to be represented by I, <strong>and</strong> ‘1’ only appears on the leading diagonal twicewhereas I appears on the leading diagonal four times in table 28.3, it is clear that no1051


GROUP THEORY1 i −1 −i1 1 i −1 −ii i −1 −i 1−1 −1 −i 1 i−i −i 1 i −11 2 4 31 1 2 4 32 2 4 3 14 4 3 1 23 3 1 2 4Table 28.5 A comparison between tables 28.4 <strong>and</strong> 28.2(b), the latter with itscolumns reordered.I A B CI I A B CA A B C IB B C I AC C I A BTable 28.6 The common structure exemplified by tables 28.4 <strong>and</strong> 28.2(b), thelatter with its columns reordered.amount of relabelling (or, equivalently, no allocation of the symbols A, B, C, amongsti, −1, −i) can bring table 28.4 into the <strong>for</strong>m of table 28.3. We conclude that the group{1,i,−1, −i} is not isomorphic to S or S ′ . An alternative way of stating the observation isto say that the group contains only one element of order 2 whilst a group correspondingto table 28.3 contains three such elements.However, if the rows <strong>and</strong> columns of table 28.2(b) – in which the identity does appeartwice on the diagonal <strong>and</strong> which there<strong>for</strong>e has the potential to be equivalent to table 28.4 –are rearranged by making the heading order 1, 2, 4, 3 then the two tables can be comparedin the <strong>for</strong>ms shown in table 28.5. They can thus be seen to have the same structure, namelythat shown in table 28.6.We there<strong>for</strong>e conclude that the group of four elements {1,i,−1, −i} under ordinary multiplicationof complex numbers is isomorphic to the group {1, 2, 3, 4} under multiplication(mod 5). ◭What we have done does not prove it, but the two tables 28.3 <strong>and</strong> 28.6 are infact the only possible tables <strong>for</strong> a group of order 4, i.e. a group containing exactlyfour elements.28.3 Non-Abelian groupsSo far, all the groups <strong>for</strong> which we have constructed multiplication tables havebeen based on some <strong>for</strong>m of arithmetic multiplication, a commutative operation,with the result that the groups have been Abelian <strong>and</strong> the tables symmetricabout the leading diagonal. We now turn to examples of groups in which somenon-commutation occurs. It should be noted, in passing, that non-commutationcannot occur throughout a group, as the identity always commutes with anyelement in its group.1052


28.3 NON-ABELIAN GROUPSAs a first example we consider again as elements of a group the two-dimensionaloperations which trans<strong>for</strong>m an equilateral triangle into itself (see the end ofsubsection 28.1.1). It has already been shown that there are six such operations:the null operation, two rotations (by 2π/3 <strong>and</strong>4π/3 about an axis perpendicularto the plane of the triangle) <strong>and</strong> three reflections in the perpendicular bisectorsof the three sides. To abbreviate we will denote these operations by symbols asfollows.(i) I is the null operation.(ii) R <strong>and</strong> R ′ are (clockwise) rotations by 2π/3 <strong>and</strong>4π/3 respectively.(iii) K, L, M are reflections in the three lines indicated in figure 28.2.Some products of the operations of the <strong>for</strong>m X • Y (where it will be recalledthat the symbol • means that the second operation X is carried out on the systemresulting from the application of the first operation Y ) are easily calculated:R • R = R ′ , R ′ • R ′ = R, R • R ′ = I = R ′ • RK • K = L • L = M • M = I.(28.12)Others, such as K • M, are more difficult, but can be found by a little thought,or by making a model triangle or drawing a sequence of diagrams such as thosefollowing.xxK • M = K == R ′xxshowing that K • M = R ′ . In the same way,M • K = M == Rx x xxshows that M • K = R, <strong>and</strong>R • L = R == Kx x x xshows that R • L = K.Proceeding in this way we can build up the complete multiplication table(table 28.7). In fact, it is not necessary to draw any more diagrams, as allremaining products can be deduced algebraically from the three found above <strong>and</strong>1053


GROUP THEORYI R R ′ K L MI I R R ′ K L MR R R ′ I M K LR ′ R ′ I R L M KK K L M I R R ′L L M K R ′ I RM M K L R R ′ ITable 28.7 The group table <strong>for</strong> the two-dimensional symmetry operations onan equilateral triangle.the more self-evident results given in (28.12). A number of things may be noticedabout this table.(i) It is not symmetric about the leading diagonal, indicating that some pairsof elements in the group do not commute.(ii) There is some symmetry within the 3×3 blocks that <strong>for</strong>m the four quartersof the table. This occurs because we have elected to put similar operationsclose to each other when choosing the order of table headings – the tworotations (or three if I is viewed as a rotation by 0π/3) are next to eachother, <strong>and</strong> the three reflections also occupy adjacent columns <strong>and</strong> rows.We will return to this later.That two groups of the same order may be isomorphic carries over to non-Abelian groups. The next two examples are each concerned with sets of sixobjects; they will be shown to <strong>for</strong>m groups that, although very different in naturefrom the rotation–reflection group just considered, are isomorphic to it.We consider first the set M of six orthogonal 2 × 2 matrices given by)) ( )−1I =( 1 00 1( ) −1 0C =0 1A =D =(−12√32− √ 32− 1 2( 12− √ 32− √ 32− 1 2)B =E =− √ 32√232− 1 2( √1 32 2√32− 1 2) (28.13)the combination law being that of ordinary matrix multiplication. Here we useitalic, rather than the sans serif used <strong>for</strong> matrices elsewhere, to emphasise thatthe matrices are group elements.Although it is tedious to do so, it can be checked that the product of anytwo of these matrices, in either order, is also in the set. However, the result isgenerally different in the two cases, as matrix multiplication is non-commutative.The matrix I clearly acts as the identity element of the set, <strong>and</strong> during the checking<strong>for</strong> closure it is found that the inverse of each matrix is contained in the set, I,C, D <strong>and</strong> E being their own inverses. The group table is shown in table 28.8.1054


28.3 NON-ABELIAN GROUPSI A B C D EI I A B C D EA A B I E C DB B I A D E CC C D E I A BD D E C B I AE E C D A B ITable 28.8 The group table, under matrix multiplication, <strong>for</strong> the set M ofsix orthogonal 2 × 2 matrices given by (28.13).The similarity to table 28.7 is striking. If {R,R ′ ,K,L,M} of that table arereplaced by {A, B, C, D, E} respectively, the two tables are identical, without eventhe need to reshuffle the rows <strong>and</strong> columns. The two groups, one of reflections<strong>and</strong> rotations of an equilateral triangle, the other of matrices, are isomorphic.Our second example of a group isomorphic to the same rotation–reflectiongroup is provided by a set of functions of an undetermined variable x. Thefunctions are as follows:f 1 (x) =x, f 2 (x) =1/(1 − x), f 3 (x) =(x − 1)/x,f 4 (x) =1/x, f 5 (x) =1− x, f 6 (x) =x/(x − 1),<strong>and</strong> the law of combination isf i (x) • f j (x) =f i (f j (x)),i.e. the function on the right acts as the argument of the function on the left toproduce a new function of x. It should be emphasised that it is the functionsthat are the elements of the group. The variable x is the ‘system’ on which theyact, <strong>and</strong> plays much the same role as the triangle does in our first example of anon-Abelian group.To show an explicit example, we calculate the product f 6 • f 3 . The productwill be the function of x obtained by evaluating y/(y − 1), when y is set equal to(x − 1)/x. Explicitly(x − 1)/xf 6 (f 3 )=(x − 1)/x − 1 =1− x = f 5(x).Thus f 6 • f 3 = f 5 . Further examples aref 2 • f 2 =11 − 1/(1 − x) = x − 1 = f 3 ,x<strong>and</strong>f 6 • f 6 =x/(x − 1)x/(x − 1) − 1 = x = f 1. (28.14)1055


GROUP THEORYThe multiplication table <strong>for</strong> this set of six functions has all the necessary propertiesto show that they <strong>for</strong>m a group. Further, if the symbols f 1 ,f 2 ,f 3 ,f 4 ,f 5 ,f 6 arereplaced by I,A,B,C,D,E respectively the table becomes identical to table 28.8.This justifies our earlier claim that this group of functions, with argument substitutionas the law of combination, is isomorphic to the group of reflections <strong>and</strong>rotations of an equilateral triangle.28.4 Permutation groupsThe operation of rearranging n distinct objects amongst themselves is called apermutation of degree n, <strong>and</strong> since many symmetry operations on physical systemscan be viewed in that light, the properties of permutations are of interest. Forexample, the symmetry operations on an equilateral triangle, to which we havealready given much attention, can be considered as the six possible rearrangementsof the marked corners of the triangle amongst three fixed points in space, muchas in the diagrams used to compute table 28.7. In the same way, the symmetryoperations on a cube can be viewed as a rearrangement of its corners amongsteight points in space, albeit with many constraints, or, with fewer complications,as a rearrangement of its body diagonals in space. The details will be left untilwe review the possible finite groups more systematically.The notations <strong>and</strong> conventions used in the literature to describe permutationsare very varied <strong>and</strong> can easily lead to confusion. We will try to avoid this by usingletters a,b,c,... (rather than numbers) <strong>for</strong> the objects that are rearranged by apermutation <strong>and</strong> by adopting, be<strong>for</strong>e long, a ‘cycle notation’ <strong>for</strong> the permutationsthemselves. It is worth emphasising that it is the permutations, i.e.theactsofrearranging, <strong>and</strong> not the objects themselves (represented by letters) that <strong>for</strong>mthe elements of permutation groups. The complete group of all permutations ofdegree n is usually denoted by S n or Σ n . The number of possible permutations ofdegree n is n!, <strong>and</strong> so this is the order of S n .Suppose the ordered set of six distinct objects {a bcdef} is rearranged bysome process into {b efadc}; then we can represent this mathematically asθ{a bcdef} = {b efadc},where θ is a permutation of degree 6. The permutation θ can be denoted by[256143],sincethefirstobject,a, is replaced by the second, b, the secondobject, b, is replaced by the fifth, e, the third by the sixth, f, etc.Theequationcan then be written more explicitly asθ{a bcdef} =[256143]{a bcdef} = {b efadc}.If φ is a second permutation, also of degree 6, then the obvious interpretation ofthe product φ • θ of the two permutations isφ • θ{a bcdef} = φ(θ{a bcdef}).1056


28.4 PERMUTATION GROUPSSuppose that φ is the permutation [4 5 3 6 2 1]; thenφ • θ{a bcdef} =[453621][256143]{a bcdef}=[453621]{b efadc}= {a dfceb}=[146352]{a bcdef}.Written in terms of the permutation notation this result is[453621][256143]=[146352].A concept that is very useful <strong>for</strong> working with permutations is that of decompositioninto cycles. The cycle notation is most easily explained by example. Forthe permutation θ given above:the 1st object, a, has been replaced by the 2nd, b;the 2nd object, b, has been replaced by the 5th, e;the 5th object, e, has been replaced by the 4th, d;the 4th object, d, has been replaced by the 1st, a.This brings us back to the beginning of a closed cycle, which is convenientlyrepresented by the notation (1 2 5 4), in which the successive replacementpositionsareenclosed,insequence,inparentheses.Thus(1254)means2nd→ 1st, 5th → 2nd, 4th → 5th, 1st → 4th. It should be noted that the objectinitially in the first listed position replaces that in the final position indicated inthe bracket – here ‘a’ is put into the fourth position by the permutation. Clearlythe cycle (5 4 1 2), or any other that involved the same numbers in the samerelative order, would have exactly the same meaning <strong>and</strong> effect. The remainingtwo objects, c <strong>and</strong> f, are interchanged by θ or, more <strong>for</strong>mally, are rearrangedaccording to a cycle of length 2, a transposition, represented by (3 6). Thus thecomplete representation (specification) of θ isθ =(1254)(36).The positions of objects that are unaltered by a permutation are either placed bythemselves in a pair of parentheses or omitted altogether. The <strong>for</strong>mer is recommendedas it helps to indicate how many objects are involved – important whenthe object in the last position is unchanged, or the permutation is the identity,which leaves all objects unaltered in position! Thus the identity permutation ofdegree 6 isI = (1)(2)(3)(4)(5)(6),though in practice it is often shortened to (1).It will be clear that the cycle representation is unique, to within the internalabsolute ordering of the numbers in each bracket as already noted, <strong>and</strong> that1057


GROUP THEORYeach number appears once <strong>and</strong> only once in the representation of any particularpermutation.The order of any permutation of degree n within the group S n canbereadofffrom the cyclic representation <strong>and</strong> is given by the lowest common multiple (LCM)of the lengths of the cycles. Thus I has order 1, as it must, <strong>and</strong> the permutationθ discussed above has order 4 (the LCM of 4 <strong>and</strong> 2).Expressed in cycle notation our second permutation φ is (3)(1 4 6)(2 5), <strong>and</strong>the product φ • θ is calculated as(3)(1 4 6)(2 5) • (1 2 5 4)(3 6){a bcdef} = (3)(1 4 6)(2 5){b efadc}= {a dfceb}= (1)(5)(2 4 3 6){a bcdef}.i.e. expressed as a relationship amongst the elements of the group of permutationsof degree 6 (not yet proved as a group, but reasonably anticipated), this resultreads(3)(1 4 6)(2 5) • (1 2 5 4)(3 6) = (1)(5)(2 4 3 6).We note, <strong>for</strong> practice, that φ has order 6 (the LCM of 1, 3, <strong>and</strong> 2) <strong>and</strong> that theproduct φ • θ has order 4.The number of elements in the group S n of all permutations of degree n isn! <strong>and</strong> clearly increases very rapidly as n increases. Fortunately, to illustrate theessential features of permutation groups it is sufficient to consider the case n =3,which involves only six elements. They are as follows (with labelling which thereader will by now recognise as anticipatory):I = (1)(2)(3) A =(123) B =(132)C = (1)(2 3) D = (3)(1 2) E = (2)(1 3)It will be noted that A <strong>and</strong> B have order 3, whilst C, D <strong>and</strong> E have order 2. Asperhaps anticipated, their combination products are exactly those correspondingto table 28.8, I, C, D <strong>and</strong> E being their own inverses. For example, putting in allsteps explicitly,D • C{a bc} = (3)(1 2) • (1)(2 3){a bc}= (3)(12){a cb}= {c ab}=(321){a bc}=(132){a bc}= B{a bc}.In brief, the six permutations belonging to S 3 <strong>for</strong>m yet another non-Abelian groupisomorphic to the rotation–reflection symmetry group of an equilateral triangle.1058


28.5 MAPPINGS BETWEEN GROUPS28.5 Mappings between groupsNow that we have available a range of groups that can be used as examples,we return to the study of more general group properties. From here on, whenthere is no ambiguity we will write the product of two elements, X • Y , simplyas XY , omitting the explicit combination symbol. We will also continue to use‘multiplication’ as a loose generic name <strong>for</strong> the combination process betweenelements of a group.If G <strong>and</strong> G ′ are two groups, we can study the effect of a mappingΦ:G → G ′of G onto G ′ .IfX is an element of G we denote its image in G ′ under the mappingΦbyX ′ =Φ(X).A technical term that we have already used is isomorphic. We will now defineit <strong>for</strong>mally. Two groups G = {X,Y,...} <strong>and</strong> G ′ = {X ′ ,Y ′ ,...} are said to beisomorphic if there is a one-to-one correspondencebetween their elements such thatX ↔ X ′ ,Y↔ Y ′ , ···XY = Z implies X ′ Y ′ = Z ′<strong>and</strong> vice versa.In other words, isomorphic groups have the same (multiplication) structure,although they may differ in the nature of their elements, combination law <strong>and</strong>notation. Clearly if groups G <strong>and</strong> G ′ are isomorphic, <strong>and</strong> G <strong>and</strong> G ′′ are isomorphic,then it follows that G ′ <strong>and</strong> G ′′ are isomorphic. We have already seen an example offour groups (of functions of x, of orthogonal matrices, of permutations <strong>and</strong> of thesymmetries of an equilateral triangle) that are isomorphic, all having table 28.8as their multiplication table.Although our main interest is in isomorphic relationships between groups, thewider question of mappings of one set of elements onto another is of someimportance, <strong>and</strong> we start with the more general notion of a homomorphism.Let G <strong>and</strong> G ′ be two groups <strong>and</strong> Φ a mapping of G → G ′ . If <strong>for</strong> every pair ofelements X <strong>and</strong> Y in G(XY ) ′ = X ′ Y ′then Φ is called a homomorphism, <strong>and</strong> G ′ is said to be a homomorphic image of G.The essential defining relationship, expressed by (XY ) ′ = X ′ Y ′ , is that thesame result is obtained whether the product of two elements is <strong>for</strong>med first <strong>and</strong>the image then taken or the images are taken first <strong>and</strong> the product then <strong>for</strong>med.1059


GROUP THEORYThree immediate consequences of the above definition are proved as follows.(i) If I is the identity of G then IX = X <strong>for</strong> all X in G. ConsequentlyX ′ =(IX) ′ = I ′ X ′ ,<strong>for</strong> all X ′ in G ′ . Thus I ′ is the identity in G ′ . In words, the identity elementof G maps into the identity element of G ′ .(ii) Further,I ′ =(XX −1 ) ′ = X ′ (X −1 ) ′ .That is, (X −1 ) ′ =(X ′ ) −1 . In words, the image of an inverse is the sameelement in G ′ as the inverse of the image.(iii) If element X in G is of order m, i.e.I = X m ,thenI ′ =(X m ) ′ =(XX m−1 ) ′ = X ′ (X m−1 ) ′ = ···= X} ′ X ′ {{ ···X}′ .m factorsIn words, the image of an element has the same order as the element.What distinguishes an isomorphism from the more general homomorphism arethe requirements that in an isomorphism:(I) different elements in G must map into different elements in G ′ (whereas ina homomorphism several elements in G may have the same image in G ′ ),that is, x ′ = y ′ must imply x = y;(II) any element in G ′ must be the image of some element in G.An immediate consequence of (I) <strong>and</strong> result (iii) <strong>for</strong> homomorphisms is thatisomorphic groups each have the same number of elements of any given order.For a general homomorphism, the set of elements of G whose image in G ′is I ′ is called the kernel of the homomorphism; this is discussed further in thenext section. In an isomorphism the kernel consists of the identity I alone. Toillustrate both this point <strong>and</strong> the general notion of a homomorphism, considera mapping between the additive group of real numbers R <strong>and</strong> the multiplicativegroup of complex numbers with unit modulus, U(1). Suppose that the mappingR→U(1) isΦ:x → e ix ;then this is a homomorphism since(x + y) ′ → e i(x+y) = e ix e iy = x ′ y ′ .However, it is not an isomorphism because many (an infinite number) of theelements of R have the same image in U(1). For example, π, 3π, 5π,... in R allhave the image −1 inU(1) <strong>and</strong>, furthermore, all elements of R of the <strong>for</strong>m 2πn,where n is an integer, map onto the identity element in U(1). The latter set <strong>for</strong>msthe kernel of the homomorphism.1060


28.6 SUBGROUPS(a)I A B C D EI I A B C D EA A B I E C DB B I A D E CC C D E I A BD D E C B I AE E C D A B I(b)I A B CI I A B CA A I C BB B C I AC C B A ITable 28.9 Reproduction of (a) table 28.8 <strong>and</strong> (b) table 28.3 with the relevantsubgroups shown in bold.For the sake of completeness, we add that a homomorphism <strong>for</strong> which (I) aboveholds is said to be a monomorphism (or an isomorphism into), whilst a homomorphism<strong>for</strong> which (II) holds is called an epimorphism (or an isomorphism onto). If,in either case, the other requirement is met as well then the monomorphism orepimorphism is also an isomorphism.Finally, if the initial <strong>and</strong> final groups are the same, G = G ′ , then the isomorphismG → G ′ is termed an automorphism.28.6 SubgroupsMore detailed inspection of tables 28.8 <strong>and</strong> 28.3 shows that not only do thecomplete tables have the properties associated with a group multiplication table(see section 28.2) but so do the upper left corners of each table taken on theirown. The relevant parts are shown in bold in the tables 28.9(a) <strong>and</strong> (b).This observation immediately prompts the notion of a subgroup. A subgroupof a group G can be <strong>for</strong>mally defined as any non-empty subset H = {H i } ofG, the elements of which themselves behave as a group under the same rule ofcombination as applies in G itself. As <strong>for</strong> all groups, the order of the subgroup isequal to the number of elements it contains; we will denote it by h or |H|.Any group G contains two trivial subgroups:(i) G itself;(ii) the set I consisting of the identity element alone.All other subgroups of G are termed proper subgroups. In a group with multiplicationtable 28.8 the elements {I,A,B} <strong>for</strong>m a proper subgroup, as do {I,A} in agroup with table 28.3 as its group table.Some groups have no proper subgroups. For example, the so-called cyclicgroups, mentioned at the end of subsection 28.1.1, have no subgroups otherthan the whole group or the identity alone. Tables 28.10(a) <strong>and</strong> (b) show themultiplication tables <strong>for</strong> two of these groups. Table 28.6 is also the group table<strong>for</strong> a cyclic group, that of order 4.1061


GROUP THEORY(a)I A BI I A BA A B IB B I A(b)I A B C DI I A B C DA A B C D IB B C D I AC C D I A BD D I A B CTable 28.10 The group tables of two cyclic groups, of orders 3 <strong>and</strong> 5. Theyhave no proper subgroups.It will be clear that <strong>for</strong> a cyclic group G repeated combination of any elementwith itself generates all other elements of G, be<strong>for</strong>e finally reproducing itself. So,<strong>for</strong> example, in table 28.10(b), starting with (say) D, repeated combination withitself produces, in turn, C, B, A, I <strong>and</strong> finally D again. As noted earlier, in anycyclic group G every element, apart from the identity, is of order g, the order ofthe group itself.The two tables shown are <strong>for</strong> groups of orders 3 <strong>and</strong> 5. It will be proved insubsection 28.7.2 that the order of any group is a multiple of the order of any ofits subgroups (Lagrange’s theorem), i.e. in our general notation, g is a multipleof h. It thus follows that a group of order p, wherep is any prime, must be cyclic<strong>and</strong> cannot have any proper subgroups. The groups <strong>for</strong> which tables 28.10(a) <strong>and</strong>(b) are the group tables are two such examples. Groups of non-prime order may(table 28.3) or may not (table 28.6) have proper subgroups.As we have seen, repeated multiplication of an element X (not the identity)by itself will generate a subgroup {X,X 2 ,X 3 ,...}. The subgroup will clearly beAbelian, <strong>and</strong> if X is of order m, i.e.X m = I, the subgroup will have m distinctmembers. If m is less than g – though, in view of Lagrange’s theorem, m mustbe a factor of g – the subgroup will be a proper subgroup. We can deduce, inpassing, that the order of any element of a group is an exact divisor of the orderof the group.Some obvious properties of the subgroups of a group G, which can be listedwithout <strong>for</strong>mal proof, are as follows.(i) The identity element of G belongs to every subgroup H.(ii) If element X belongs to a subgroup H, so does X −1 .(iii) The set of elements in G that belong to every subgroup of G themselves<strong>for</strong>m a subgroup, though this may consist of the identity alone.Properties of subgroups that need more explicit proof are given in the followingsections, though some need the development of new concepts be<strong>for</strong>e theycan be established. However, we can begin with a theorem, applicable to allhomomorphisms, not just isomorphisms, that requires no new concepts.Let Φ : G → G ′ be a homomorphism of G into G ′ ;then1062


28.7 SUBDIVIDING A GROUP(i) the set of elements H ′ in G ′ that are images of the elements of G <strong>for</strong>ms asubgroup of G ′ ;(ii) the set of elements K in G that are mapped onto the identity I ′ in G ′ <strong>for</strong>msa subgroup of G.As indicated in the previous section, the subgroup K is called the kernel of thehomomorphism.To prove (i), suppose Z <strong>and</strong> W belong to H ′ , with Z = X ′ <strong>and</strong> W = Y ′ ,whereX <strong>and</strong> Y belong to G. Then<strong>and</strong> there<strong>for</strong>e belongs to H ′ ,<strong>and</strong>ZW = X ′ Y ′ =(XY ) ′Z −1 =(X ′ ) −1 =(X −1 ) ′<strong>and</strong> there<strong>for</strong>e belongs to H ′ . These two results, together with the fact that I ′belongs to H ′ , are enough to establish result (i).To prove (ii), suppose X <strong>and</strong> Y belong to K; then(XY ) ′ = X ′ Y ′ = I ′ I ′ = I ′(closure),I ′ =(XX −1 ) ′ = X ′ (X −1 ) ′ = I ′ (X −1 ) ′ =(X −1 ) ′<strong>and</strong> there<strong>for</strong>e X −1 belongs to K. These two results, together with the fact that Ibelongs to K, are enough to establish (ii). An illustration of this result is providedby the mapping Φ of R→U(1) considered in the previous section. Its kernelconsists of the set of real numbers of the <strong>for</strong>m 2πn, wheren is an integer; it <strong>for</strong>msa subgroup of R, the additive group of real numbers.In fact the kernel K of a homomorphism is a normal subgroup of G. Thedefining property of such a subgroup is that <strong>for</strong> every element X in G <strong>and</strong> everyelement Y in the subgroup, XY X −1 belongs to the subgroup. This property iseasily verified <strong>for</strong> the kernel K, since(XY X −1 ) ′ = X ′ Y ′ (X −1 ) ′ = X ′ I ′ (X −1 ) ′ = X ′ (X −1 ) ′ = I ′ .Anticipating the discussion of subsection 28.7.2, the cosets of a normal subgroupthemselves <strong>for</strong>m a group (see exercise 28.16).28.7 Subdividing a groupWe have already noted, when looking at the (arbitrary) order of headings in agroup table, that some choices appear to make the table more orderly than doothers. In the following subsections we will identify ways in which the elementsof a group can be divided up into sets with the property that the members of anyone set are more like the other members of the set, in some particular regard,1063


GROUP THEORYthan they are like any element that does not belong to the set. We will find thatthese divisions will be such that the group is partitioned, i.e. the elements will bedivided into sets in such a way that each element of the group belongs to one,<strong>and</strong> only one, such set.We note in passing that the subgroups of a group do not <strong>for</strong>m such a partition,not least because the identity element is in every subgroup, rather than being inprecisely one. In other words, despite the nomenclature, a group is not simply theaggregate of its proper subgroups.28.7.1 Equivalence relations <strong>and</strong> classesWe now specify in a more mathematical manner what it means <strong>for</strong> two elementsof a group to be ‘more like’ one another than like a third element, as mentionedin section 28.2. Our introduction will apply to any set, whether a group or not,but our main interest will ultimately be in two particular applications to groups.We start with the <strong>for</strong>mal definition of an equivalence relation.An equivalence relation on a set S is a relationship X ∼ Y , between twoelements X <strong>and</strong> Y belonging to S, in which the definition of the symbol ∼ mustsatisfy the requirements of(i) reflexivity, X ∼ X;(ii) symmetry, X ∼ Y implies Y ∼ X;(iii) transitivity, X ∼ Y <strong>and</strong> Y ∼ Z imply X ∼ Z.Any particular two elements either satisfy or do not satisfy the relationship.The general notion of an equivalence relation is very straight<strong>for</strong>ward, <strong>and</strong>the requirements on the symbol ∼ seem undem<strong>and</strong>ing; but not all relationshipsqualify. As an example within the topic of groups, if it meant ‘has the sameorder as’ then clearly all the requirements would be satisfied. However, if it meant‘commutes with’ then it would not be an equivalence relation, since although Acommutes with I, <strong>and</strong>I commutes with C, this does not necessarily imply that Acommutes with C, as is obvious from table 28.8.It may be shown that an equivalence relation on S divides up S into classes C isuch that:(i) X <strong>and</strong> Y belong to the same class if, <strong>and</strong> only if, X ∼ Y ;(ii) every element W of S belongs to exactly one class.This may be shown as follows. Let X belong to S, <strong>and</strong> define the subset S X ofS to be the set of all elements U of S such that X ∼ U. Clearly by reflexivityX belongs to S X . Suppose first that X ∼ Y , <strong>and</strong> let Z be any element of S Y .Then Y ∼ Z, <strong>and</strong> hence by transitivity X ∼ Z, which means that Z belongs toS X . Conversely, since the symmetry law gives Y ∼ X, ifZ belongs to S X then1064


28.7 SUBDIVIDING A GROUPthis implies that Z belongs to S Y . These two results together mean that the twosubsets S X <strong>and</strong> S Y have the same members <strong>and</strong> hence are equal.Now suppose that S X equals S Y .SinceY belongs to S Y it also belongs to S X<strong>and</strong> hence X ∼ Y . This completes the proof of (i), once the distinct subsets oftype S X are identified as the classes C i . Statement (ii) is an immediate corollary,the class in question being identified as S W .The most important property of an equivalence relation is as follows.Two different subsets S X <strong>and</strong> S Y can have no element in common, <strong>and</strong> the collectionof all the classes C i is a ‘partition’ of S, i.e. every element in S belongs to one, <strong>and</strong>only one, of the classes.To prove this, suppose S X <strong>and</strong> S Y have an element Z in common; then X ∼ Z<strong>and</strong> Y ∼ Z <strong>and</strong> so by the symmetry <strong>and</strong> transitivity laws X ∼ Y .Bytheabovetheorem this implies S X equals S Y . But this contradicts the fact that S X <strong>and</strong> S Yare different subsets. Hence S X <strong>and</strong> S Y can have no element in common.Finally, if the elements of S are used in turn to define subsets <strong>and</strong> hence classesin S, every element U is in the subset S U that is either a class already found orconstitutes a new one. It follows that the classes exhaust S, i.e.everyelementisin some class.Having established the general properties of equivalence relations, we now turnto two specific examples of such relationships, in which the general set S has themore specialised properties of a group G <strong>and</strong> the equivalence relation ∼ is chosenin such a way that the relatively transparent general results <strong>for</strong> equivalencerelations can be used to derive powerful, but less obvious, results about theproperties of groups.28.7.2 Congruence <strong>and</strong> cosetsAs the first application of equivalence relations we now prove Lagrange’s theoremwhich is stated as follows.Lagrange’s theorem. If G is a finite group of order g <strong>and</strong> H is a subgroup of G o<strong>for</strong>der h then g is a multiple of h.We take as the definition of ∼ that, given X <strong>and</strong> Y belonging to G, X ∼ Y ifX −1 Y belongs to H. This is the same as saying that Y = XH i <strong>for</strong> some elementH i belonging to H; technically X <strong>and</strong> Y are said to be left-congruent with respectto H.This defines an equivalence relation, since it has the following properties.(i) Reflexivity: X ∼ X, sinceX −1 X = I <strong>and</strong> I belongs to any subgroup.(ii) Symmetry: X ∼ Y implies that X −1 Y belongs to H <strong>and</strong> so, there<strong>for</strong>e, doesits inverse, since H is a group. But (X −1 Y ) −1 = Y −1 X <strong>and</strong>, as this belongsto H, it follows that Y ∼ X.1065


GROUP THEORY(iii) Transitivity: X ∼ Y <strong>and</strong> Y ∼ Z imply that X −1 Y <strong>and</strong> Y −1 Z belong to H<strong>and</strong> so, there<strong>for</strong>e, does their product (X −1 Y )(Y −1 Z)=X −1 Z,fromwhichit follows that X ∼ Z.With ∼ proved as an equivalence relation, we can immediately deduce that itdivides G into disjoint (non-overlapping) classes. For this particular equivalencerelation the classes are called the left cosets of H. Thus each element of G is inone <strong>and</strong> only one left coset of H. The left coset containing any particular X isusually written XH, <strong>and</strong> denotes the set of elements of the <strong>for</strong>m XH i (one ofwhich is X itself since H contains the identity element); it must contain h differentelements, since if it did not, <strong>and</strong> two elements were equal,XH i = XH j ,we could deduce that H i = H j <strong>and</strong> that H contained fewer than h elements.From our general results about equivalence relations it now follows that theleft cosets of H are a ‘partition’ of G into a number of sets each containing hmembers. Since there are g members of G <strong>and</strong> each must be in just one of thesets, it follows that g is a multiple of h. This concludes the proof of Lagrange’stheorem.The number of left cosets of H in G is known as the index of H in G <strong>and</strong> iswritten [G : H]; numerically the index = g/h. For the record we note that, <strong>for</strong>the trivial subgroup I, which contains only the identity element, [G : I] =g <strong>and</strong>that, <strong>for</strong> a subgroup J of subgroup H, [G : H][H : J ]=[G : J ].The validity of Lagrange’s theorem was established above using the far-reachingproperties of equivalence relations. However, <strong>for</strong> this specific purpose there is amore direct <strong>and</strong> self-contained proof, which we now give.Let X be some particular element of a finite group G of order g, <strong>and</strong>H be asubgroup of G of order h, with typical element Y i . Consider the set of elementsXH ≡{XY 1 ,XY 2 ,...,XY h }.This set contains h distinct elements, since if any two were equal, i.e. XY i = XY jwith i ≠ j, this would contradict the cancellation law. As we have already seen,the set is called a left coset of H.We now prove three simple results.• Two cosets are either disjoint or identical. Suppose cosets X 1 H <strong>and</strong> X 2 H havean element in common, i.e. X 1 Y 1 = X 2 Y 2 <strong>for</strong> some Y 1 ,Y 2 in H. ThenX 1 =X 2 Y 2 Y1 −1 , <strong>and</strong> since Y 1 <strong>and</strong> Y 2 both belong to H so does Y 2 Y1 −1 ; thus X 1belongs to the left coset X 2 H. Similarly X 2 belongs to the left coset X 1 H.Consequently, either the two cosets are identical or it was wrong to assumethat they have an element in common.1066


28.7 SUBDIVIDING A GROUP• Two cosets X 1 H <strong>and</strong> X 2 H are identical if, <strong>and</strong> only if, X2 −1X1 belongs to H. IfX2 −1X1 belongs to H then X 1 = X 2 Y i <strong>for</strong> some i, <strong>and</strong>X 1 H = X 2 Y i H = X 2 H,since by the permutation law Y i H = H. Thus the two cosets are identical.Conversely, suppose X 1 H = X 2 H.ThenX2 −1X1H = H. But one element ofH (on the left of the equation) is I; thus X2 −1X1 must also be an element of H(on the right). This proves the stated result.• Every element of G is in some left coset XH. This follows trivially since Hcontains I, <strong>and</strong> so the element X i is in the coset X i H.The final step in establishing Lagrange’s theorem is, as previously, to note thateach coset contains h elements, that the cosets are disjoint <strong>and</strong> that every one ofthe g elements in G appears in one <strong>and</strong> only one distinct coset. It follows thatg = kh <strong>for</strong> some integer k.As noted earlier, Lagrange’s theorem justifies our statement that any group o<strong>for</strong>der p, wherep is prime, must be cyclic <strong>and</strong> cannot have any proper subgroups:since any subgroup must have an order that divides p, this can only be 1 or p,corresponding to the two trivial subgroups I <strong>and</strong> the whole group.It may be helpful to see an example worked through explicitly, <strong>and</strong> we againuse the same six-element group.◮Find the left cosets of the proper subgroup H of the group G that has table 28.8 as itsmultiplication table.The subgroup consists of the set of elements H = {I,A,B}. We note in passing that it hasorder 3, which, as required by Lagrange’s theorem, is a divisor of 6, the order of G. Asinall cases, H itself provides the first (left) coset, <strong>for</strong>mally the cosetIH = {II,IA,IB} = {I,A,B}.We continue by choosing an element not already selected, C say, <strong>and</strong> <strong>for</strong>mCH = {CI,CA,CB} = {C,D,E}.These two cosets of H exhaust G, <strong>and</strong> are there<strong>for</strong>e the only cosets, the index of H in Gbeing equal to 2.This completes the example, but it is useful to demonstrate that it would not havemattered if we had taken D, say, instead of I to <strong>for</strong>m a first cosetDH = {DI, DA, DB} = {D, E, C},<strong>and</strong> then, from previously unselected elements, picked B, say:BH = {BI,BA,BB} = {B,I,A}.The same two cosets would have resulted. ◭It will be noticed that the cosets are the same groupings of the elementsof G which we earlier noted as being the choice of adjacent column <strong>and</strong> rowheadings that give the multiplication table its ‘neatest’ appearance. Furthermore,1067


GROUP THEORYif H is a normal subgroup of G then its (left) cosets themselves <strong>for</strong>m a group (seeexercise 28.16).28.7.3 Conjugates <strong>and</strong> classesOur second example of an equivalence relation is concerned with those elementsX <strong>and</strong> Y of a group G that can be connected by a trans<strong>for</strong>mation of the <strong>for</strong>mY = G −1i XG i ,whereG i is an (appropriate) element of G. Thus X ∼ Y if thereexists an element G i of G such that Y = G −1i XG i . Different pairs of elements X<strong>and</strong> Y will, in general, require different group elements G i . Elements connectedin this way are said to be conjugates.We first need to establish that this does indeed define an equivalence relation,as follows.(i) Reflexivity: X ∼ X, sinceX = I −1 XI <strong>and</strong> I belongs to the group.(ii) Symmetry: X ∼ Y implies Y = G −1i XG i <strong>and</strong> there<strong>for</strong>e X =(G −1i ) −1 YG −1i .Since G i belongs to G so does G −1i , <strong>and</strong> it follows that Y ∼ X.(iii) Transitivity: X ∼ Y <strong>and</strong> Y ∼ Z imply Y = G −1i XG i <strong>and</strong> Z = G −1j YG j<strong>and</strong> there<strong>for</strong>e Z = G −1j G −1i XG i G j =(G i G j ) −1 X(G i G j ). Since G i <strong>and</strong> G jbelong to G so does G i G j , from which it follows that X ∼ Z.These results establish conjugacy as an equivalence relation <strong>and</strong> hence showthat it divides G into classes, two elements being in the same class if, <strong>and</strong> only if,they are conjugate.Immediate corollaries are:(i) If Z is in the class containing I thenZ = G −1i IG i = G −1i G i = I.Thus, since any conjugate of I canbeshowntobeI, the identity must bein a class by itself.(ii) If X is in a class by itself thenY = G −1i XG imust imply that Y = X. ButX = G i G −1i XG i G −1i<strong>for</strong> any G i ,<strong>and</strong>soX = G i (G −1i XG i )G −1i = G i YG −1i = G i XG −1i ,i.e. XG i = G i X <strong>for</strong> all G i .Thus commutation with all elements of the group is a necessary (<strong>and</strong>sufficient) condition <strong>for</strong> any particular group element to be in a class byitself. In an Abelian group each element is in a class by itself.1068


28.7 SUBDIVIDING A GROUP(iii) In any group G the set S of elements in classes by themselves is an Abeliansubgroup (known as the centre of G). We have shown that I belongs to S,<strong>and</strong> so if, further, XG i = G i X <strong>and</strong> YG i = G i Y <strong>for</strong> all G i belonging to Gthen:(a) (XY )G i = XG i Y = G i (XY ), i.e. the closure of S, <strong>and</strong>(b) XG i = G i X implies X −1 G i = G i X −1 , i.e. the inverse of X belongsto S.Hence S is a group, <strong>and</strong> clearly Abelian.Yet again <strong>for</strong> illustration purposes, we use the six-element group that hastable 28.8 as its group table.◮Find the conjugacy classes of the group G having table 28.8 as its multiplication table.As always, I is in a class by itself, <strong>and</strong> we need consider it no further.Consider next the results of <strong>for</strong>ming X −1 AX, asX runs through the elements of G.I −1 AI A −1 AA B −1 AB C −1 AC D −1 AD E −1 AE= IA = IA = AI = CE = DC = ED= A = A = A = B = B = BOnly A <strong>and</strong> B are generated. It is clear that {A, B} is one of the conjugacy classes of G.This can be verified by <strong>for</strong>ming all elements X −1 BX; again only A <strong>and</strong> B appear.We now need to pick an element not in the two classes already found. Suppose wepick C. Justas<strong>for</strong>A, we compute X −1 CX, asX runs through the elements of G. Thecalculations can be done directly using the table <strong>and</strong> give the following:X : I A B C D EX −1 CX : C E D C E DThus C, D <strong>and</strong> E belong to the same class. The group is now exhausted, <strong>and</strong> so the threeconjugacy classes are{I}, {A, B}, {C,D,E}. ◭In the case of this small <strong>and</strong> simple, but non-Abelian, group, only the identityis in a class by itself (i.e. only I commutes with all other elements). It is also theonly member of the centre of the group.Other areas from which examples of conjugacy classes can be taken includepermutations <strong>and</strong> rotations. Two permutations can only be (but are not necessarily)in the same class if their cycle specifications have the same structure.For example, in S 5 the permutations (1 3 5)(2)(4) <strong>and</strong> (2 5 3)(1)(4) could be inthe same class as each other but not in the class that contains (1 5)(2 4)(3). Anexample of permutations with the same cycle structure yet in different conjugacyclasses is given in exercise 29. 10.In the case of the continuous rotation group, rotations by the same angle θabout any two axes labelled i <strong>and</strong> j are in the same class, because the groupcontains a rotation that takes the first axis into the second. Without going into1069


GROUP THEORYmathematical details, a rotation about axis i can be represented by the operatorR i (θ), <strong>and</strong> the two rotations are connected by a relationship of the <strong>for</strong>mR j (θ) =φ −1ij R i (θ)φ ij ,in which φ ij is the member of the full continuous rotation group that takes axisi into axis j.28.8 Exercises28.1 For each of the following sets, determine whether they <strong>for</strong>m a group under the operationindicated (where it is relevant you may assume that matrix multiplicationis associative):(a) the integers (mod 10) under addition;(b) the integers (mod 10) under multiplication;(c) the integers 1, 2, 3, 4, 5, 6 under multiplication (mod 7);(d) the integers 1, 2, 3, 4, 5 under multiplication (mod 6);(e)all matrices of the <strong>for</strong>m(a a− b0 bwhere a <strong>and</strong> b are integers (mod 5) <strong>and</strong> a ≠0≠ b, under matrix multiplication;(f) those elements of the set in (e) that are of order 1 or 2 (taken together);(g) all matrices of the <strong>for</strong>m⎛⎝ 1 a 0 1 00⎞⎠ ,b c 1where a, b, c are integers, under matrix multiplication.28.2 Which of the following relationships between X <strong>and</strong> Y are equivalence relations?Give a proof of your conclusions in each case:(a) X <strong>and</strong> Y are integers <strong>and</strong> X − Y is odd;(b) X <strong>and</strong> Y are integers <strong>and</strong> X − Y is even;(c) X <strong>and</strong> Y are people <strong>and</strong> have the same postcode;(d) X <strong>and</strong> Y are people <strong>and</strong> have a parent in common;(e) X <strong>and</strong> Y are people <strong>and</strong> have the same mother;(f) X <strong>and</strong> Y are n×n matrices satisfying Y = PXQ,whereP <strong>and</strong> Q are elementsof a group G of n × n matrices.28.3 Define a binary operation • on the set of real numbers byx • y = x + y + rxy,where r is a non-zero real number. Show that the operation • is associative.Prove that x • y = −r −1 if, <strong>and</strong> only if, x = −r −1 or y = −r −1 . Hence provethat the set of all real numbers excluding −r −1 <strong>for</strong>ms a group under the operation•.1070),


28.8 EXERCISES28.4 Prove that the relationship X ∼ Y , defined by X ∼ Y if Y can be expressed inthe <strong>for</strong>mY = aX + bcX + d ,with a, b, c <strong>and</strong> d as integers, is an equivalence relation on the set of real numbersR. Identify the class that contains the real number 1.28.5 The following is a ‘proof’ that reflexivity is an unnecessary axiom <strong>for</strong> an equivalencerelation.Because of symmetry X ∼ Y implies Y ∼ X. Then by transitivity X ∼ Y <strong>and</strong>Y ∼ X imply X ∼ X. Thus symmetry <strong>and</strong> transitivity imply reflexivity, whichthere<strong>for</strong>e need not be separately required.Demonstrate the flaw in this proof using the set consisting of all real numbers plusthe number i. Show by investigating the following specific cases that, whether ornot reflexivity actually holds, it cannot be deduced from symmetry <strong>and</strong> transitivityalone.(a) X ∼ Y if X + Y is real.(b) X ∼ Y if XY is real.28.6 Prove that the set M of matricesA =(a b0 cwhere a, b, c are integers (mod 5) <strong>and</strong> a ≠0≠ c, <strong>for</strong>m a non-Abelian groupunder matrix multiplication.Show that the subset containing elements of M that are of order 1 or 2 donot <strong>for</strong>m a proper subgroup of M,(a) using Lagrange’s theorem,(b) by direct demonstration that the set is not closed.28.7 S is the set of all 2 × 2 matrices of the <strong>for</strong>m( )w xA =, where wz − xy =1.y zShow that S is a group under matrix multiplication. Which element(s) have order2? Prove that an element A has order 3 if w + z +1=0.28.8 Show that, under matrix multiplication, matrices of the <strong>for</strong>m( )a0 + aM(a 0 , a) =1 i −a 2 + a 3 i,a 2 + a 3 i a 0 − a 1 iwhere a 0 <strong>and</strong> the components of column matrix a =(a 1 a 2 a 3 ) T are realnumbers satisfying a 2 0 + |a|2 = 1, constitute a group. Deduce that, under thetrans<strong>for</strong>mation z → Mz, wherez is any column matrix, |z| 2 is invariant.28.9 If A is a group in which every element other than the identity, I, hasorder2,prove that A is Abelian. Hence show that if X <strong>and</strong> Y are distinct elements of A,neither being equal to the identity, then the set {I,X,Y ,XY } <strong>for</strong>ms a subgroupof A.Deduce that if B is a group of order 2p, withp a prime greater than 2, then Bmust contain an element of order p.28.10 The group of rotations (excluding reflections <strong>and</strong> inversions) in three dimensionsthat take a cube into itself is known as the group 432 (or O in the usual chemicalnotation). Show by each of the following methods that this group has 24 elements.1071),


GROUP THEORY(a) Identify the distinct relevant axes <strong>and</strong> count the number of qualifying rotationsabout each.(b) The orientation of the cube is determined if the directions of two of its bodydiagonals are given. Consider the number of distinct ways in which one bodydiagonal can be chosen to be ‘vertical’, say, <strong>and</strong> a second diagonal made tolie along a particular direction.28.11 Identify the eight symmetry operations on a square. Show that they <strong>for</strong>m agroup D 4 (known to crystallographers as 4mm <strong>and</strong>tochemistsasC 4v ) having oneelement of order 1, five of order 2 <strong>and</strong> two of order 4. Find its proper subgroups<strong>and</strong> the corresponding cosets.28.12 If A <strong>and</strong> B are two groups, then their direct product, A × B, isdefinedtobethe set of ordered pairs (X,Y ), with X an element of A, Y an element of B<strong>and</strong> multiplication given by (X,Y )(X ′ ,Y ′ )=(XX ′ ,YY ′ ). Prove that A × B is agroup.Denote the cyclic group of order n by C n <strong>and</strong> the symmetry group of a regularn-sided figure (an n-gon) by D n – thus D 3 is the symmetry group of an equilateraltriangle, as discussed in the text.(a) By considering the orders of each of their elements, show (i) that C 2 × C 3 isisomorphic to C 6 , <strong>and</strong> (ii) that C 2 × D 3 is isomorphic to D 6 .(b) Are any of D 4 , C 8 , C 2 × C 4 , C 2 × C 2 × C 2 isomorphic?28.13 Find the group G generated under matrix multiplication by the matrices( ) ( )0 10 iA =, B =.1 0i 0Determine its proper subgroups, <strong>and</strong> verify <strong>for</strong> each of them that its cosetsexhaust G.28.14 Show that if p is prime then the set of rational number pairs (a, b), excluding(0, 0), with multiplication defined by(a, b) • (c, d) =(e, f), where (a + b √ p)(c + d √ p)=e + f √ p,<strong>for</strong>ms an Abelian group. Show further that the mapping (a, b) → (a, −b) isanautomorphism.28.15 Consider the following mappings between a permutation group <strong>and</strong> a cyclicgroup.(a) Denote by A n the subset of the permutation group S n that contains all theeven permutations. Show that A n is a subgroup of S n .(b) List the elements of S 3 in cycle notation <strong>and</strong> identify the subgroup A 3 .(c) For each element X of S 3 ,letp(X) =1ifX belongs to A 3 <strong>and</strong> p(X) =−1 ifitdoes not. Denote by C 2 the multiplicative cyclic group of order 2. Determinethe images of each of the elements of S 3 <strong>for</strong> the following four mappings:Φ 1 : S 3 → C 2 X → p(X),Φ 2 : S 3 → C 2 X →−p(X),Φ 3 : S 3 → A 3 X → X 2 ,Φ 4 : S 3 → S 3 X → X 3 .(d) For each mapping, determine whether the kernel K is a subgroup of S 3 <strong>and</strong>,if so, whether the mapping is a homomorphism.28.16 For the group G with multiplication table 28.8 <strong>and</strong> proper subgroup H = {I,A,B},denote the coset {I,A,B} by C 1 <strong>and</strong> the coset {C,D,E} by C 2 .Formthesetofall possible products of a member of C 1 with itself, <strong>and</strong> denote this by C 1 C 1 .1072


28.8 EXERCISESSimilarly compute C 2 C 2 , C 1 C 2 <strong>and</strong> C 2 C 1 . Show that each product coset is equal toC 1 or to C 2 ,<strong>and</strong>thata2× 2 multiplication table can be <strong>for</strong>med, demonstratingthat C 1 <strong>and</strong> C 2 are themselves the elements of a group of order 2. A subgrouplike H whose cosets themselves <strong>for</strong>m a group is a normal subgroup.28.17 The group of all non-singular n × n matrices is known as the general lineargroup GL(n) <strong>and</strong> that with only real elements as GL(n, R). If R ∗ denotes themultiplicative group of non-zero real numbers, prove that the mapping Φ :GL(n, R) → R ∗ , defined by Φ(M) =detM, is a homomorphism.Show that the kernel K of Φ is a subgroup of GL(n, R). Determine its cosets<strong>and</strong> show that they themselves <strong>for</strong>m a group.28.18 The group of reflection–rotation symmetries of a square is known as D 4 ;letX be one of its elements. Consider a mapping Φ : D 4 → S 4 , the permutationgroup on four objects, defined by Φ(X) = the permutation induced by X onthe set {x, y, d, d ′ }, where x <strong>and</strong> y are the two principal axes, <strong>and</strong> d <strong>and</strong> d ′the two principal diagonals, of the square. For example, if R is a rotation byπ/2, Φ(R) = (12)(34). Show that D 4 is mapped onto a subgroup of S 4 <strong>and</strong>, byconstructing the multiplication tables <strong>for</strong> D 4 <strong>and</strong> the subgroup, prove that themapping is a homomorphism.28.19 Given that matrix M is a member of the multiplicative group GL(3, R), determine,<strong>for</strong> each of the following additional constraints on M (applied separately), whetherthe subset satisfying the constraint is a subgroup of GL(3, R):(a) M T = M;(b) M T M = I;(c) |M| =1;(d) M ij =0<strong>for</strong>j>i<strong>and</strong> M ii ≠0.28.20 The elements of the quaternion group, Q, aretheset{1, −1,i,−i, j, −j,k,−k},with i 2 = j 2 = k 2 = −1, ij = k <strong>and</strong> its cyclic permutations, <strong>and</strong> ji = −k <strong>and</strong>its cyclic permutations. Find the proper subgroups of Q <strong>and</strong> the correspondingcosets. Show that the subgroup of order 2 is a normal subgroup, but that theother subgroups are not. Show that Q cannot be isomorphic to the group 4mm(C 4v ) considered in exercise 28.11.28.21 Show that D 4 , the group of symmetries of a square, has two isomorphic subgroupsof order 4. Show further that there exists a two-to-one homomorphism from thequaternion group Q, of exercise 28.20, onto one (<strong>and</strong> hence either) of these twosubgroups, <strong>and</strong> determine its kernel.28.22 Show that the matrices⎛M(θ, x, y) = ⎝cos θ − sin θ xsin θ cos θ y0 0 1where 0 ≤ θ


GROUP THEORYm 1 (π)m 2 (π)m 3 (π)m 4 (π)Figure 28.3 The notation <strong>for</strong> exercise 28.11.28.9 Hints <strong>and</strong> answers28.1 § (a) Yes, (b) no, there is no inverse <strong>for</strong> 2, (c) yes, (d) no, 2 × 3 is not in the set,(e) yes, (f) yes, they <strong>for</strong>m a subgroup of order 4, [1, 0; 0, 1] [4, 0; 0, 4] [1, 2; 0, 4][4, 3; 0, 1], (g) yes.28.3 x • (y • z) =x + y + z + r(xy + xz + yz)+r 2 xyz =(x • y) • z. Show that assumingx • y = −r −1 leads to (rx +1)(ry +1)=0. The inverse of x is x −1 = −x/(1 + rx);show that this is not equal to −r −1 .28.5 (a) Consider both X = i <strong>and</strong> X ≠ i. Here, i ≁ i. (b) In this case i ∼ i, but theconclusion cannot be deduced from the other axioms. In both cases i is in a classby itself <strong>and</strong> no Y , as used in the false proof, can be found.28.7 † Use |AB| = |A||B| =1×1 = 1 to prove closure. The inverse has w ↔ z, x ↔−x,y ↔−y, giving |A −1 | = 1, i.e. it is in the set. The only element of order 2 is −I;A 2 can be simplified to [−(w +1), −x; −y, −(z + 1)].28.9 If XY = Z, show that Y = XZ <strong>and</strong> X = ZY,then<strong>for</strong>mYX. Note that theelements of B can only have orders 1, 2 or p. Suppose they all have order 1 or2; then using the earlier result, whilst noting that 4 does not divide 2p, leadstoa contradiction.28.11 Using the notation indicated in figure 28.3, R being a rotation of π/2 about anaxis perpendicular to the square, we have: I has order 1; R 2 , m 1 , m 2 , m 3 , m 4 haveorder 2; R, R 3 have order 4.subgroup {I,R,R 2 ,R 3 } has cosets {I,R,R 2 ,R 3 }, {m 1 ,m 2 ,m 3 ,m 4 };subgroup {I,R 2 ,m 1 ,m 2 } has cosets {I,R 2 ,m 1 ,m 2 }, {R,R 3 ,m 3 ,m 4 };subgroup {I,R 2 ,m 3 ,m 4 } has cosets {I,R 2 ,m 3 ,m 4 }, {R,R 3 ,m 1 ,m 2 };subgroup {I,R 2 } has cosets {I,R 2 }, {R,R 3 }, {m 1 ,m 2 }, {m 3 ,m 4 };subgroup {I,m 1 } has cosets {I,m 1 }, {R,m 3 }, {R 2 ,m 2 }, {R 3 ,m 4 };subgroup {I,m 2 } has cosets {I,m 2 }, {R,m 4 }, {R 2 ,m 1 }, {R 3 ,m 3 };subgroup {I,m 3 } has cosets {I,m 3 }, {R,m 2 }, {R 2 ,m 4 }, {R 3 ,m 1 };subgroup {I,m 4 } has cosets {I,m 4 }, {R,m 1 }, {R 2 ,m 3 }, {R 3 ,m 2 }.28.13 G = {I, A, B, B 2 , B 3 , AB, AB 2 , AB 3 }. The proper subgroups are as follows:{I, A}, {I, B 2 }, {I, AB 2 }, {I, B, B 2 , B 3 }, {I, B 2 , AB, AB 3 }.28.15 (b) A 3 = {(1), (123), (132)}.(d) For Φ 1 , K = {(1), (123), (132)} is a subgroup.For Φ 2 , K = {(23), (13), (12)} is not a subgroup because it has no identity element.For Φ 3 , K = {(1), (23), (13), (12)} is not a subgroup because it is not closed.§ Where matrix elements are given as a list, the convention used is [row 1; row 2; ...], individualentries in each row being separated by commas.1074


28.9 HINTS AND ANSWERS≠For Φ 4 , K = {(1), (123), (132)} is a subgroup.Only Φ 1 is a homomorphism; Φ 4 fails because, <strong>for</strong> example, [(23)(13)] ′(23) ′ (13) ′ .28.17 Recall that, <strong>for</strong> any pair of matrices P <strong>and</strong> Q, |PQ| = |P||Q|. K is the set of allmatrices with unit determinant. The cosets of K are the sets of matrices whosedeterminants are equal; K itself is the identity in the group of cosets.28.19 (a) No, because the set is not closed, (b) yes, (c) yes, (d) yes.28.21 Each subgroup contains the identity, a rotation by π, <strong>and</strong>tworeflections.Thehomomorphism is ±1 → I, ±i → R 2 , ±j → m x , ±k → m y with kernel {1, −1}.28.23 There are 10 elements in all: I, rotations R i (i =1, 4) <strong>and</strong> reflections m j (j =1, 5).(a) There are five proper subgroups of order 2, {I,m j } <strong>and</strong> one proper subgroupof order 5, {I,R,R 2 ,R 3 ,R 4 }.(b) Four conjugacy classes, {I}, {R,R 4 }, {R 2 ,R 3 }, {m 1 ,m 2 ,m 3 ,m 4 ,m 5 }.1075


29Representation theoryAs indicated at the start of the previous chapter, significant conclusions canoften be drawn about a physical system simply from the study of its symmetryproperties. That chapter was devoted to setting up a <strong>for</strong>mal mathematical basis,group theory, with which to describe <strong>and</strong> classify such properties; the currentchapter shows how to implement the consequences of the resulting classifications<strong>and</strong> obtain concrete physical conclusions about the system under study. Theconnection between the two chapters is akin to that between working withcoordinate-free vectors, each denoted by a single symbol, <strong>and</strong> working with acoordinate system in which the same vectors are expressed in terms of components.The ‘coordinate systems’ that we will choose will be ones that are expressed interms of matrices; it will be clear that ordinary numbers would not be sufficient,as they make no provision <strong>for</strong> any non-commutation amongst the elementsof a group. Thus, in this chapter the group elements will be represented bymatrices that have the same commutation relations as the members of the group,whatever the group’s original nature (symmetry operations, functional <strong>for</strong>ms,matrices, permutations, etc.). For some abstract groups it is difficult to give awritten description of the elements <strong>and</strong> their properties without recourse to suchrepresentations. Most of our applications will be concerned with representationsof the groups that consist of the symmetry operations on molecules containingtwo or more identical atoms.Firstly, in section 29.1, we use an elementary example to demonstrate the kindof conclusions that can be reached by arguing purely on symmetry grounds. Thenin sections 29.2–29.10 we develop the <strong>for</strong>mal side of representation theory <strong>and</strong>establish general procedures <strong>and</strong> results. Finally, these are used in section 29.11to tackle a variety of problems drawn from across the physical sciences.1076


29.1 DIPOLE MOMENTS OF MOLECULESABA ′ B ′(a) HCl (b) CO 2 (c) O 3Figure 29.1 Three molecules, (a) hydrogen chloride, (b) carbon dioxide <strong>and</strong>(c) ozone, <strong>for</strong> which symmetry considerations impose varying degrees ofconstraint on their possible electric dipole moments.29.1 Dipole moments of moleculesSome simple consequences of symmetry can be demonstrated by consideringwhether a permanent electric dipole moment can exist in any particular molecule;three simple molecules, hydrogen chloride, carbon dioxide <strong>and</strong> ozone, are illustratedin figure 29.1. Even if a molecule is electrically neutral, an electric dipolemoment will exist in it if the centres of gravity of the positive charges (due toprotons in the atomic nuclei) <strong>and</strong> of the negative charges (due to the electrons)do not coincide.For hydrogen chloride there is no reason why they should coincide; indeed, thenormal picture of the binding mechanism in this molecule is that the electron fromthe hydrogen atom moves its average position from that of its proton nucleus tosomewhere between the hydrogen <strong>and</strong> chlorine nuclei. There is no compensatingmovement of positive charge, <strong>and</strong> a net dipole moment is to be expected – <strong>and</strong>is found experimentally.For the linear molecule carbon dioxide it seems obvious that it cannot havea dipole moment, because of its symmetry. Putting this rather more rigorously,we note that any rotation about the long axis of the molecule leaves it totallyunchanged; consequently, any component of a permanent electric dipole perpendicularto that axis must be zero (a non-zero component would rotate althoughno physical change had taken place in the molecule). That only leaves the possibilityof a component parallel to the axis. However, a rotation of π radiansabout the axis AA ′ shown in figure 29.1(b) carries the molecule into itself, asdoes a reflection in a plane through the carbon atom <strong>and</strong> perpendicular to themolecular axis (i.e. one with its normal parallel to the axis). In both cases the twooxygen atoms change places but, as they are identical, the molecule is indistinguishablefrom the original. Either ‘symmetry operation’ would reverse the signof any dipole component directed parallel to the molecular axis; this can only becompatible with the indistinguishability of the original <strong>and</strong> final systems if theparallel component is zero. Thus on symmetry grounds carbon dioxide cannothave a permanent electric dipole moment.1077


REPRESENTATION THEORYFinally, <strong>for</strong> ozone, which is angular rather than linear, symmetry does notplace such tight constraints. A dipole-moment component parallel to the axisBB ′ (figure 29.1(c)) is possible, since there is no symmetry operation that reversesthe component in that direction <strong>and</strong> at the same time carries the molecule intoan indistinguishable copy of itself. However, a dipole moment perpendicular toBB ′ is not possible, since a rotation of π about BB ′ would both reverse anysuch component <strong>and</strong> carry the ozone molecule into itself – two contradictoryconclusions unless the component is zero.In summary, symmetry requirements appear in the <strong>for</strong>m that some or allcomponents of permanent electric dipoles in molecules are <strong>for</strong>bidden; they donot show that the other components do exist, only that they may. The greaterthe symmetry of the molecule, the tighter the restrictions on potentially non-zerocomponents of its dipole moment.In section 23.11 other, more complicated, physical situations will be analysedusing results derived from representation theory. In anticipation of these results,<strong>and</strong> since it may help the reader to underst<strong>and</strong> where the developments in thenext nine sections are leading, we make here a broad, powerful, but rather <strong>for</strong>mal,statement as follows.If a physical system is such that after the application of particular rotations orreflections (or a combination of the two) the final system is indistinguishable fromthe original system then its behaviour, <strong>and</strong> hence the functions that describe itsbehaviour, must have the corresponding property of invariance when subjected tothe same rotations <strong>and</strong> reflections.29.2 Choosing an appropriate <strong>for</strong>malismAs mentioned in the introduction to this chapter, the elements of a finite groupG can be represented by matrices; this is done in the following way. A suitablecolumn matrix u, known as a basis vector, § is chosen <strong>and</strong> is written in terms ofits components u i ,thebasis functions, asu =(u 1 u 2 ··· u n ) T .Theu i may be ofa variety of natures, e.g. numbers, coordinates, functions or even a set of labels,though <strong>for</strong> any one basis vector they will all be of the same kind.Once chosen, the basis vector can be used to generate an n-dimensional representationof the group as follows. An element X of the group is selected <strong>and</strong>its effect on each basis function u i is determined. If the action of X on u 1 is toproduce u ′ 1 , etc. then the set of equationsu ′ i = Xu i (29.1)§ This usage of the term basis vector is not exactly the same as that introduced in subsection 8.1.1.1078


29.2 CHOOSING AN APPROPRIATE FORMALISMgenerates a new column matrix u ′ =(u ′ 1 u′ 2 ··· u′ n) T . Having established u <strong>and</strong> u ′we can determine the n × n matrix, M(X) say, that connects them byu ′ = M(X)u. (29.2)It may seem natural to use the matrix M(X) so generated as the representativematrix of the element X; in fact, because we have already chosen the conventionwhereby Z = XY implies that the effect of applying element Z is the same as thatof first applying Y <strong>and</strong> then applying X to the result, one further step has to betaken. So that the representative matrices D(X) may follow the same convention,i.e.D(Z) =D(X)D(Y ),<strong>and</strong> at the same time respect the normal rules of matrix multiplication, it isnecessary to take the transpose of M(X) as the representative matrix D(X).Explicitly,<strong>and</strong> (29.2) becomesD(X) =M T (X) (29.3)u ′ = D T (X)u. (29.4)Thus the procedure <strong>for</strong> determining the matrix D(X) that represents the groupelement X in a representation based on basis vector u is summarised by equations(29.1)–(29.4). §This procedure is then repeated <strong>for</strong> each element X of the group, <strong>and</strong> theresulting set of n × n matrices D = {D(X)} is said to be the n-dimensionalrepresentation of G having u as its basis. The need to take the transpose of eachmatrix M(X) is not of any fundamental significance, since the only thing thatreally matters is whether the matrices D(X) have the appropriate multiplicationproperties – <strong>and</strong>, as defined, they do.In cases in which the basis functions are labels, the actions of the groupelements are such as to cause rearrangements of the labels. Correspondingly thematrices D(X) contain only ‘1’s <strong>and</strong> ‘0’s as entries; each row <strong>and</strong> each columncontains a single ‘1’.§ An alternative procedure in which a row vector is used as the basis vector is possible. Definingequations of the <strong>for</strong>m u T X = u T D(X) are used, <strong>and</strong> no additional transpositions are needed todefine the representative matrices. However, row-matrix equations are cumbersome to write out<strong>and</strong> in all other parts of this book we have adopted the convention of writing operators (here thegroup element) to the left of the object on which they operate (here the basis vector).1079


REPRESENTATION THEORY◮For the group S 3 of permutations on three objects, which has group multiplication table28.8 on p. 1055, with (in cycle notation)I = (1)(2)(3), A =(123), B =(132C = (1)(2 3), D = (3)(1 2), E = (2)(1 3),use as the components of a basis vector the ordered letter tripletsu 1 = {PQR}, u 2 = {QRP}, u 3 = {RPQ},u 4 = {PRQ}, u 5 = {QPR}, u 6 = {RQP}.Generate a six-dimensional representation D = {D(X)} of the group <strong>and</strong> confirm that therepresentative matrices multiply according to table 28.8, e.g.D(C)D(B) =D(E).It is immediate that the identity permutation I = (1)(2)(3) leaves all u i unchanged, i.e.u ′ i = u i <strong>for</strong> all i. The representative matrix D(I) is thus I 6 ,the6× 6 unit matrix.We next take X as the permutation A = (1 2 3) <strong>and</strong>, using (29.1), let it act on each ofthe components of the basis vector:u ′ 1 = Au 1 =(123){PQR} = {QRP} = u 2u ′ 2 = Au 2 =(123){QRP} = {RPQ} = u 3..u ′ 6 = Au 6 =(123){RQP} = {QPR} = u 5 .The matrix M(A) has to be such that u ′ = M(A)u (here dots replace zeros to aid readability):⎛u ′ =⎜⎝u 2u 3u 1u 6u 4u 5⎞ ⎛=⎟ ⎜⎠ ⎝· 1 · · · ·· · 1 · · ·1 · · · · ·· · · · · 1· · · 1 · ·· · · · 1 ·⎞ ⎛⎟ ⎜⎠ ⎝u 1u 2u 3u 4u 5u 6⎞≡ M(A)u.⎟⎠D(A) isthenequaltoM T (A).The other D(X) are determined in a similar way. In general, ifXu i = u j ,then [M(X)] ij = 1, leading to [D(X)] ji =1<strong>and</strong>[D(X)] jk =0<strong>for</strong>k ≠ i. For example,Cu 3 = (1)(23){RPQ} = {RQP} = u 6implies that [D(C)] 63 =1<strong>and</strong>[D(C)] 6k =0<strong>for</strong>k =1, 2, 4, 5, 6. When calculated in full⎛⎞⎛⎞· · · 1 · ·· 1 · · · ·· · · · 1 ·· · 1 · · ·· · · · · 11 · · · · ·D(C) =⎜ 1 · · · · ·, D(B) =⎟⎜ · · · · · 1,⎟⎝ · 1 · · · · ⎠⎝ · · · 1 · · ⎠· · 1 · · ·· · · · 1 ·⎛⎞· · · · · 1· · · 1 · ·· · · · 1 ·D(E) =⎜ · 1 · · · ·,⎟⎝ · · 1 · · · ⎠1 · · · · ·1080


29.2 CHOOSING AN APPROPRIATE FORMALISMP1P3P3R322Q11 2RQ R(a) (b) (c)QFigure 29.2 Diagram (a) shows the definition of the basis vector, (b) showsthe effect of applying a clockwise rotation of 2π/3 <strong>and</strong> (c) shows the effect ofapplying a reflection in the mirror axis through Q.from which it can be verified that D(C)D(B) =D(E). ◭Whilst a representation obtained in this way necessarily has the same dimensionas the order of the group it represents, there are, in general, square matrices ofboth smaller <strong>and</strong> larger dimensions that can be used to represent the group,though their existence may be less obvious.One possibility that arises when the group elements are symmetry operationson an object whose position <strong>and</strong> orientation can be referred to a spacecoordinate system is called the natural representation. In it the representativematrices D(X) describe, in terms of a fixed coordinate system, what happensto a coordinate system that moves with the object when X is applied. Thereis usually some redundancy of the coordinates used in this type of representation,since interparticle distances are fixed <strong>and</strong> fewer than 3N coordinates,where N is the number of identical particles, are needed to specify uniquelythe object’s position <strong>and</strong> orientation. Subsection 29.11.1 gives an example thatillustrates both the advantages <strong>and</strong> disadvantages of the natural representation.We continue here with an example of a natural representation that has no suchredundancy.◮Use the fact that the group considered in the previous worked example is isomorphic tothe group of two-dimensional symmetry operations on an equilateral triangle to generate athree-dimensional representation of the group.Label the triangle’s corners as 1, 2, 3 <strong>and</strong> three fixed points in space as P, Q, R, so thatinitially corner 1 lies at point P, 2 lies at point Q, <strong>and</strong> 3 at point R. We take P, Q, R asthe components of the basis vector.In figure 29.2, (a) shows the initial configuration <strong>and</strong> also, <strong>for</strong>mally, the result of applyingthe identity I to the triangle; it is there<strong>for</strong>e described by the basis vector, (P Q R) T .Diagram (b) shows the the effect of a clockwise rotation by 2π/3, corresponding toelement A in the previous example; the new column matrix is (Q R P) T .Diagram (c) shows the effect of a typical mirror reflection – the one that leaves thecorner at point Q unchanged (element D in table 28.8 <strong>and</strong> the previous example); the newcolumn matrix is now (R Q P) T .In similar fashion it can be concluded that the column matrix corresponding to elementB, rotationby4π/3, is (R P Q) T , <strong>and</strong> that the other two reflections C <strong>and</strong> E result in1081


REPRESENTATION THEORYcolumn matrices (P R Q) T <strong>and</strong> (Q P R) T respectively. The <strong>for</strong>ms of the representativematrices M nat (X), (29.2), are now determined by equations such as, <strong>for</strong> element E,⎛⎝ Q ⎞ ⎛P ⎠ = ⎝ 0 1 0⎞ ⎛1 0 0 ⎠ ⎝ P ⎞Q ⎠R 0 0 1 Rimplying that⎛D nat (E) = ⎝ 0 1 0⎞T⎛1 0 0 ⎠ = ⎝ 0 1 0⎞1 0 0 ⎠ .0 0 1 0 0 1In this way the complete representation is obtained as⎛D nat (I) = ⎝ 1 0 0⎞⎛0 1 0 ⎠ , D nat (A) = ⎝ 0 0 1⎞⎛1 0 0 ⎠ , D nat (B) = ⎝ 0 1 0⎞0 0 1 ⎠ ,0 0 10 1 01 0 0⎛D nat (C) = ⎝ 1 0 0⎞⎛0 0 1 ⎠ , D nat (D) = ⎝ 0 0 1⎞⎛0 1 0 ⎠ , D nat (E) = ⎝ 0 1 0⎞1 0 0 ⎠ .0 1 01 0 00 0 1It should be emphasised that although the group contains six elements this representationis three-dimensional. ◭We will concentrate on matrix representations of finite groups, particularlyrotation <strong>and</strong> reflection groups (the so-called crystal point groups). The generalideas carry over to infinite groups, such as the continuous rotation groups, but ina book such as this, which aims to cover many areas of applicable mathematics,some topics can only be mentioned <strong>and</strong> not explored. We now give the <strong>for</strong>maldefinition of a representation.Definition. A representation D = {D(X)} of a group G is an assignment of a nonsingularsquare n × n matrix D(X) to each element X belonging to G, such that(i) D(I) =I n , the unit n × n matrix,(ii) D(X)D(Y )=D(XY ) <strong>for</strong> any two elements X <strong>and</strong> Y belonging to G, i.e.thematrices multiply in the same way as the group elements they represent.As mentioned previously, a representation by n × n matrices is said to be ann-dimensional representation of G. The dimension n is not to be confused withg, the order of the group, which gives the number of matrices needed in therepresentation, though they might not all be different.A consequence of the two defining conditions <strong>for</strong> a representation is that thematrix associated with the inverse of X is the inverse of the matrix associatedwith X. This follows immediately from setting Y = X −1 in (ii):D(X)D(X −1 )=D(XX −1 )=D(I) =I n ;henceD(X −1 )=[D(X)] −1 .1082


29.2 CHOOSING AN APPROPRIATE FORMALISMAs an example, the four-element Abelian group that consists of the set {1,i,−1, −i}under ordinary multiplication has a two-dimensional representation based on thecolumn matrix (1 i) T :( )( )1 00 −1D(1) =, D(i) =,D(−1) =0 1( −1 00 −1), D(−i) =1 0( 0 1−1 0The reader should check that D(i)D(−i) =D(1), D(i)D(i) =D(−1) etc., i.e. thatthe matrices do have exactly the same multiplication properties as the elementsof the group. Having done so, the reader may also wonder why anybody wouldbother with the representative matrices, when the original elements are so muchsimpler to h<strong>and</strong>le! As we will see later, once some general properties of matrixrepresentations have been established, the analysis of large groups, both Abelian<strong>and</strong> non-Abelian, can be reduced to routine, almost cookbook, procedures.An n-dimensional representation of G is a homomorphism of G into the set ofinvertible n × n matrices (i.e. n × n matrices that have inverses or, equivalently,have non-zero determinants); this set is usually known as the general linear group<strong>and</strong> denoted by GL(n). In general the same matrix may represent more than oneelement of G; if, however, all the matrices representing the elements of G aredifferent then the representation is said to be faithful, <strong>and</strong> the homomorphismbecomes an isomorphism onto a subgroup of GL(n).A trivial but important representation is D(X) =I n <strong>for</strong> all elements X of G.Clearly both of the defining relationships are satisfied, <strong>and</strong> there is no restrictionon the value of n. However, such a representation is not a faithful one.To sum up, in the context of a rotation–reflection group, the transposes ofthe set of n × n matrices D(X) thatmakeuparepresentationD may be thoughtof as describing what happens to an n-component basis vector of coordinates,(x y ···) T , or of functions, (Ψ 1 Ψ 2 ···) T ,theΨ i themselves being functionsof coordinates, when the group operation X is carried out on each of thecoordinates or functions. For example, to return to the symmetry operationson an equilateral triangle, the clockwise rotation by 2π/3, R, carries the threedimensionalbasis vector (x y z) T into the column matrix⎛⎜⎝− 1 2 x + √ 32 y− √ 32 x − 1 2 yzwhilst the two-dimensional basis vector of functions (r 2 3z 2 − r 2 ) T is unaltered,as neither r nor z is changed by the rotation. The fact that z is unchanged byany of the operations of the group shows that the components x, y, z actuallydivide (i.e. are ‘reducible’, to anticipate a more <strong>for</strong>mal description) into two sets:1083⎞⎟⎠).


REPRESENTATION THEORYone comprises z, which is unchanged by any of the operations, <strong>and</strong> the othercomprises x, y, which change as a pair into linear combinations of themselves.This is an important observation to which we return in section 29.4.29.3 Equivalent representationsIf D is an n-dimensional representation of a group G, <strong>and</strong>Q is any fixed invertiblen × n matrix (|Q| ≠ 0), then the set of matrices defined by the similaritytrans<strong>for</strong>mationD Q (X) =Q −1 D(X)Q (29.5)also <strong>for</strong>ms a representation D Q of G, said to be equivalent to D. Wecanseefromacomparison with the definition in section 29.2 that they do <strong>for</strong>m a representation:(i) D Q (I) =Q −1 D(I)Q = Q −1 I n Q = I n ,(ii) D Q (X)D Q (Y )=Q −1 D(X)QQ −1 D(Y )Q = Q −1 D(X)D(Y )Q= Q −1 D(XY )Q = D Q (XY ).Since we can always trans<strong>for</strong>m between equivalent representations using a nonsingularmatrix Q, we will consider such representations to be one <strong>and</strong> the same.Despite the similarity of words <strong>and</strong> manipulations to those of subsection 28.7.1,that two representations are equivalent does not constitute an ‘equivalence relation’– <strong>for</strong> example, the reflexive property does not hold <strong>for</strong> a general fixedmatrix Q. However, if Q were not fixed, but simply restricted to belonging toa set of matrices that themselves <strong>for</strong>m a group, then (29.5) would constitute anequivalence relation.The general invertible matrix Q that appears in the definition (29.5) of equivalentmatrices describes changes arising from a change in the coordinate system(i.e. in the set of basis functions). As be<strong>for</strong>e, suppose that the effect of an operationX on the basis functions is expressed by the action of M(X) (whichisequalto D T (X)) on the corresponding basis vector:u ′ = M(X)u = D T (X)u. (29.6)A change of basis would be given by u Q = Qu <strong>and</strong> u ′ Q = Qu′ , <strong>and</strong> we may writeThis is of the same <strong>for</strong>m as (29.6), i.e.u ′ Q = Qu ′ = QM(X)u = QD T (X)Q −1 u Q . (29.7)u ′ Q = D T Q T(X)u Q, (29.8)where D Q T(X) =(Q T ) −1 D(X)Q T is related to D(X) by a similarity trans<strong>for</strong>mation.Thus D Q T(X) represents the same linear trans<strong>for</strong>mation as D(X), but with1084


29.3 EQUIVALENT REPRESENTATIONSrespect to a new basis vector u Q ; this supports our contention that representationsconnected by similarity trans<strong>for</strong>mations should be considered as the samerepresentation.◮For the four-element Abelian group consisting of the set {1,i,−1, −i} under ordinarymultiplication, discussed near the end of section 29.2, change the basis vector from u =(1 i) T to u Q =(3− i 2i − 5) T . Find the real trans<strong>for</strong>mation matrix Q. Show that thetrans<strong>for</strong>med representative matrix <strong>for</strong> element i, D Q T(i), isgivenby( )17 −29D Q T(i) =10 −17<strong>and</strong> verify that D T Q T (i)u Q = iu Q .Firstly, we solve the matrix equation( )3 − i=2i − 5( )(a b 1c d i),with a, b, c, d real. This gives Q <strong>and</strong> hence Q −1 as( )( )3 −12 1Q =, Q−5 2−1 =.5 3Following (29.7) we now find the transpose of D Q T(i) as( )( )( ) ( )3 −1 0 1 2 1 17 10QD T (i)Q −1 ==−5 2 −1 0 5 3 −29 −17<strong>and</strong> hence D Q T(i) is as stated. Finally,(17 10D T Q T(i)u Q =−29 −17= i(3 − i2i − 5)( ) ( )3 − i 1+3i=2i − 5 −2 − 5i)= iu Q ,as required. ◭Although we will not prove it, it can be shown that any finite representationof a finite group of linear trans<strong>for</strong>mations that preserve spatial length (or, inquantum mechanics, preserve the magnitude of a wavefunction) is equivalent to1085


REPRESENTATION THEORYa representation in which all the matrices are unitary (see chapter 8) <strong>and</strong> so fromnow on we will consider only unitary representations.29.4 Reducibility of a representationWe have seen already that it is possible to have more than one representationof any particular group. For example, the group {1,i,−1, −i} under ordinarymultiplication has been shown to have a set of 2 × 2 matrices, <strong>and</strong> a set of fourunit n × n matrices I n , as two of its possible representations.Consider two or more representations, D (1) , D (2) , ...,D (N) , which may beof different dimensions, of a group G. Now combine the matrices D (1) (X),D (2) (X), ...,D (N) (X) that correspond to element X of G into a larger blockdiagonalmatrix:D (1) (X )0D(X ) =D (2) (X )(29.9). . .0D (N) (X )Then D = {D(X)} is the matrix representation of the group obtained by combiningthe basis vectors of D (1) , D (2) , ...,D (N) into one larger basis vector. If, knowingly orunknowingly, we had started with this larger basis vector <strong>and</strong> found the matricesof the representation D to have the <strong>for</strong>m shown in (29.9), or to have a <strong>for</strong>mthat can be trans<strong>for</strong>med into this by a similarity trans<strong>for</strong>mation (29.5) (using,of course, the same matrix Q <strong>for</strong> each of the matrices D(X)) then we would saythat D is reducible <strong>and</strong> that each matrix D(X) can be written as the direct sum ofsmaller representations:D(X) =D (1) (X) ⊕ D (2) (X) ⊕ ···⊕ D (N) (X).It may be that some or all of the matrices D (1) (X), D (2) (X), ...,D (N) themselvescan be further reduced – i.e. written in block diagonal <strong>for</strong>m. For example,suppose that the representation D (1) , say, has a basis vector (x y z) T ; then, <strong>for</strong>the symmetry group of an equilateral triangle, whilst x <strong>and</strong> y are mixed together<strong>for</strong> at least one of the operations X, z is never changed. In this case the 3 × 3representative matrix D (1) (X) can itself be written in block diagonal <strong>for</strong>m as a1086


29.4 REDUCIBILITY OF A REPRESENTATION2 × 2matrix<strong>and</strong>a1× 1 matrix. The direct-sum matrix D(X) can now be writtena bc d10D(X ) =D (2) (X )(29.10). . .0D (N) (X )but the first two blocks can be reduced no further.When all the other representations D (2) (X), ... have been similarly treated,what remains is said to be irreducible <strong>and</strong> has the characteristic of being blockdiagonal, with blocks that individually cannot be reduced further. The blocks areknown as the irreducible representations of G, often abbreviated to the irreps ofG, <strong>and</strong> we denote them by ˆD (i) . They <strong>for</strong>m the building blocks of representationtheory, <strong>and</strong> it is their properties that are used to analyse any given physicalsituation which is invariant under the operations that <strong>for</strong>m the elements of G.Any representation can be written as a linear combination of irreps.If, however, the initial choice u of basis vector <strong>for</strong> the representation D isarbitrary, as it is in general, then it is unlikely that the matrices D(X) willassume obviously block diagonal <strong>for</strong>ms (it should be noted, though, that sincethe matrices are square, even a matrix with non-zero entries only in the extremetop right <strong>and</strong> bottom left positions is technically block diagonal). In general, itwill be possible to reduce them to block diagonal matrices with more than oneblock; this reduction corresponds to a trans<strong>for</strong>mation Q to a new basis vectoru Q , as described in section 29.3.In any particular representation D, each constituent irrep ˆD (i) may appear anynumber of times, or not at all, subject to the obvious restriction that the sum ofall the irrep dimensions must add up to the dimension of D itself. Let us say thatˆD (i) appears m i times. The general expansion of D is then writtenD = m 1 ˆD (1) ⊕ m 2 ˆD (2) ⊕ ···⊕ m N ˆD (N) , (29.11)where if G is finite so is N.This is such an important result that we shall now restate the situation insomewhat different language. When the set of matrices that <strong>for</strong>ms a representation1087


REPRESENTATION THEORYof a particular group of symmetry operations has been brought to irreducible<strong>for</strong>m, the implications are as follows.(i) Those components of the basis vector that correspond to rows in therepresentation matrices with a single-entry block, i.e. a 1 × 1block,areunchanged by the operations of the group. Such a coordinate or functionis said to trans<strong>for</strong>m according to a one-dimensional irrep of G. Intheexample given in (29.10), that the entry on the third row <strong>for</strong>ms a 1 × 1block implies that the third entry in the basis vector (x y z ···) T ,namely z, is invariant under the two-dimensional symmetry operations onan equilateral triangle in the xy-plane.(ii) If, in any of the g matrices of the representation, the largest-sized blocklocated on the row or column corresponding to a particular coordinate(or function) in the basis vector is n × n, then that coordinate (or function)is mixed by the symmetry operations with n − 1 others <strong>and</strong> is said totrans<strong>for</strong>m according to an n-dimensional irrep of G. Thus in the matrix(29.10), x is the first entry in the complete basis vector; the first row ofthe matrix contains two non-zero entries, as does the first column, <strong>and</strong> sox is part of a two-component basis vector whose components are mixedby the symmetry operations of G. The other component is y.The result (29.11) may also be <strong>for</strong>mulated in terms of the more abstract notionof vector spaces (chapter 8). The set of g matrices that <strong>for</strong>ms an n-dimensionalrepresentation D of the group G can be thought of as acting on column matricescorresponding to vectors in an n-dimensional vector space V spanned by the basisfunctions of the representation. If there exists a proper subspace W of V ,suchthat if a vector whose column matrix is w belongs to W then the vector whosecolumn matrix is D(X)w also belongs to W , <strong>for</strong> all X belonging to G, thenitfollows that D is reducible. We say that the subspace W is invariant under theactions of the elements of G. With D unitary, the orthogonal complement W ⊥ ofW ,i.e.thevectorspaceV remaining when the subspace W has been removed, isalso invariant, <strong>and</strong> all the matrices D(X) split into two blocks acting separatelyon W <strong>and</strong> W ⊥ .BothW <strong>and</strong> W ⊥ may contain further invariant subspaces, inwhich case the matrices will be split still further.As a concrete example of this approach, consider in plane polar coordinatesρ, φ the effect of rotations about the polar axis on the infinite-dimensional vectorspace V of all functions of φ that satisfy the Dirichlet conditions <strong>for</strong> expansionas a Fourier series (see section 12.1). We take as our basis functions the set{sin mφ, cos mφ} <strong>for</strong> integer values m =0, 1, 2,...; this is an infinite-dimensionalrepresentation (n = ∞) <strong>and</strong>, since a rotation about the polar axis can be throughany angle α (0 ≤ α


29.4 REDUCIBILITY OF A REPRESENTATIONNow, <strong>for</strong> some k, consider a vector w in the space W k spanned by {sin kφ, cos kφ},say w = a sin kφ + b cos kφ. Under a rotation by α about the polar axis, a sin kφbecomes a sin k(φ + α), which can be written as a cos kα sin kφ + a sin kα cos kφ, i.eas a linear combination of sin kφ <strong>and</strong> cos kφ; similarly cos kφ becomes anotherlinear combination of the same two functions. The newly generated vector w ′ ,whose column matrix w ′ is given by w ′ = D(α)w, there<strong>for</strong>e belongs to W k <strong>for</strong>any α <strong>and</strong> we can conclude that W k is an invariant irreducible two-dimensionalsubspace of V . It follows that D(α) is reducible <strong>and</strong> that, since the result holds<strong>for</strong> every k, in its reduced <strong>for</strong>m D(α) has an infinite series of identical 2 × 2 blockson its leading diagonal; each block will have the <strong>for</strong>m( )cos α − sin α.sin α cos αWe note that the particular case k = 0 is special, in that then sin kφ = 0 <strong>and</strong>cos kφ = 1, <strong>for</strong> all φ; consequently the first 2×2 block in D(α) is reducible further<strong>and</strong> becomes two single-entry blocks.A second illustration of the connection between the behaviour of vector spacesunder the actions of the elements of a group <strong>and</strong> the <strong>for</strong>m of the matrix representationof the group is provided by the vector space spanned by the sphericalharmonics Y lm (θ, φ). This contains subspaces, corresponding to the differentvalues of l, that are invariant under the actions of the elements of the full threedimensionalrotation group; the corresponding matrices are block-diagonal, <strong>and</strong>those entries that correspond to the part of the basis containing Y lm (θ, φ) <strong>for</strong>ma(2l +1)× (2l + 1) block.To illustrate further the irreps of a group, we return again to the group G oftwo-dimensional rotation <strong>and</strong> reflection symmetries of an equilateral triangle, orequivalently the permutation group S 3 ; this may be shown, using the methods ofsection 29.7 below, to have three irreps. Firstly, we have already seen that the setM of six orthogonal 2 × 2 matrices given in section (28.3), equation (28.13), isisomorphic to G. These matrices there<strong>for</strong>e <strong>for</strong>m not only a representation of G,but a faithful one. It should be noticed that, although G contains six elements,the matrices are only 2 × 2. However, they contain no invariant 1 × 1 sub-block(which <strong>for</strong> 2 × 2 matrices would require them all to be diagonal) <strong>and</strong> neither canall the matrices be made block-diagonal by the same similarity trans<strong>for</strong>mation;they there<strong>for</strong>e <strong>for</strong>m a two-dimensional irrep of G.Secondly, as previously noted, every group has one (unfaithful) irrep in whichevery element is represented by the 1 × 1 matrix I 1 , or, more simply, 1.Thirdly an (unfaithful) irrep of G is given by assignment of the one-dimensionalset of six ‘matrices’ {1, 1, 1, −1, −1, −1} to the symmetry operations {I,R,R ′ ,K,L, M} respectively, or to the group elements {I,A,B,C,D,E} respectively; seesection 28.3. In terms of the permutation group S 3 , 1 corresponds to evenpermutations <strong>and</strong> −1 to odd permutations, ‘odd’ or ‘even’ referring to the number1089


REPRESENTATION THEORYof simple pair interchanges to which a permutation is equivalent. That theseassignments are in accord with the group multiplication table 28.8 should bechecked.Thus the three irreps of the group G (i.e. the group 3m or C 3v or S 3 ), are, usingthe conventional notation A 1 ,A 2 , E (see section 29.8), as follows:where( )1 0M I =, M A =0 1(−1)0M C =0 1, M D =ElementI A B C D EA 1 1 1 1 1 1 1(29.12)E M I M A M B M C M D M EIrrep A 2 1 1 1 −1 −1 −1(−12√32− √ 32− 1 2(12− √ 32− √ 32− 1 2)(−12− √ )32, M B = √3,2− 1 2) ( √ )1 32 2, M E = .√32− 1 229.5 The orthogonality theorem <strong>for</strong> irreducible representationsWe come now to the central theorem of representation theory, a theorem thatjustifies the relatively routine application of certain procedures to determinethe restrictions that are inherent in physical systems that have some degree ofrotational or reflection symmetry. The development of the theorem is long <strong>and</strong>quite complex when presented in its entirety, <strong>and</strong> the reader will have to referelsewhere <strong>for</strong> the proof. §The theorem states that, in a certain sense, the irreps of a group G are asorthogonal as possible, as follows. If, <strong>for</strong> each irrep, the elements in any onepositionineachoftheg matricesareusedtomakeupg-component columnmatrices then(i) any two such column matrices coming from different irreps are orthogonal;(ii) any two such column matrices coming from different positions in thematrices of the same irrep are orthogonal.This orthogonality is in addition to the irreps’ being in the <strong>for</strong>m of orthogonal(unitary) matrices <strong>and</strong> thus each comprising mutually orthogonal rows <strong>and</strong>columns.§ See, e.g., H. F. Jones, Groups, Representations <strong>and</strong> <strong>Physics</strong> (Bristol: Institute of <strong>Physics</strong>, 1998); J.F. Cornwell, Group Theory in <strong>Physics</strong>, vol 2 (London: Academic Press, 1984); J-P. Serre, LinearRepresentations of Finite Groups (New York: Springer, 1977).1090


29.5 THE ORTHOGONALITY THEOREM FOR IRREDUCIBLE REPRESENTATIONSMore mathematically, if we denote the entry in the ith row <strong>and</strong> jth column of amatrix D(X) by[D(X)] ij ,<strong>and</strong> ˆD (λ) <strong>and</strong> ˆD (µ) are two irreps of G having dimensionsn λ <strong>and</strong> n µ respectively, then∑ [ ]ˆD (λ) ∗ [(X) ˆD (X)](µ) = g δ ik δ jl δ λµ . (29.13)ijkl nXλThis rather <strong>for</strong>bidding-looking equation needs some further explanation.Firstly, the asterisk indicates that the complex conjugate should be taken ifnecessary, though all our representations so far have involved only real matrixelements. Each Kronecker delta function on the right-h<strong>and</strong> side has the value 1if its two subscripts are equal <strong>and</strong> has the value 0 otherwise. Thus the right-h<strong>and</strong>side is only non-zero if i = k, j = l <strong>and</strong> λ = µ, all at the same time.Secondly, the summation over the group elements X means that g contributionshave to be added together, each contribution being a product of entries drawnfrom the representative matrices in the two irreps ˆD (λ) = { ˆD (λ) (X)} <strong>and</strong> ˆD (µ) ={ ˆD (µ) (X)}. Theg contributions arise as X runs over the g elements of G.Thus, putting these remarks together, the summation will produce zero if either(i) the matrix elements are not taken from exactly the same position in everymatrix, including cases in which it is not possible to do so because theirreps ˆD (λ) <strong>and</strong> ˆD (µ) have different dimensions, or(ii) even if ˆD (λ) <strong>and</strong> ˆD (µ) do have the same dimensions <strong>and</strong> the matrix elementsare from the same positions in every matrix, they are different irreps, i.e.λ ≠ µ.Some numerical illustrations based on the irreps A 1 ,A 2 <strong>and</strong> E of the group 3m(or C 3v or S 3 ) will probably provide the clearest explanation (see (29.12)).(a) Take i = j = k = l = 1, with ˆD (λ) =A 1 <strong>and</strong> ˆD (µ) =A 2 . Equation (29.13)then reads1(1) + 1(1) + 1(1) + 1(−1) + 1(−1) + 1(−1) = 0,as expected, since λ ≠ µ.(b) Take (i, j) as(1, 2) <strong>and</strong> (k, l) as(2, 2), corresponding to different matrixpositions within the same irrep ˆD (λ) = ˆD (µ) = E. Substituting in (29.13)gives(0(1) +− √ 32) (−12)+( √32) (− ) (12 +0(1)+− √ 32) (−12)+( √32) (−12)=0.(c) Take (i, j) as(1, 2), <strong>and</strong> (k, l) as(1, 2), corresponding to the same matrixpositions within the same irrep ˆD (λ) = ˆD (µ) = E. Substituting in (29.13)gives(0(0)+− √ 32)(− √ 32)+( √32)( √32) (+0(0)+− √ 32)(− √ 32)+( √32)( √32)= 6 2 .1091


REPRESENTATION THEORY(d) No explicit calculation is needed to see that if i = j = k = l = 1, withˆD (λ) = ˆD (µ) =A 1 (or A 2 ), then each term in the sum is either 1 2 or (−1) 2<strong>and</strong> the total is 6, as predicted by the right-h<strong>and</strong> side of (29.13) since g =6<strong>and</strong> n λ =1.29.6 CharactersThe actual matrices of general representations <strong>and</strong> irreps are cumbersome towork with, <strong>and</strong> they are not unique since there is always the freedom to changethe coordinate system, i.e. the components of the basis vector (see section 29.3),<strong>and</strong> hence the entries in the matrices. However, one thing that does not change<strong>for</strong> a matrix under such an equivalence (similarity) trans<strong>for</strong>mation – i.e. undera change of basis – is the trace of the matrix. This was shown in chapter 8,but is repeated here. The trace of a matrix A is the sum of its diagonal elements,n∑Tr A =or, using the summation convention (section 26.1), simply A ii . Under a similaritytrans<strong>for</strong>mation, again using the summation convention,i=1A ii[D Q (X)] ii =[Q −1 ] ij [D(X)] jk [Q] ki=[D(X)] jk [Q] ki [Q −1 ] ij=[D(X)] jk [I] kj=[D(X)] jj ,showing that the traces of equivalent matrices are equal.This fact can be used to greatly simplify work with representations, though withsome partial loss of the in<strong>for</strong>mation content of the full matrices. For example,using trace values alone it is not possible to distinguish between the two groupsknown as 4mm <strong>and</strong> ¯42m, orasC 4v <strong>and</strong> D 2d respectively, even though the twogroups are not isomorphic. To make use of these simplifications we now definethe characters of a representation.Definition. The characters χ(D) of a representation D of a group G are defined asthe traces of the matrices D(X), one <strong>for</strong> each element X of G.At this stage there will be g characters, but, as we noted in subsection 28.7.3,elements A, B of G in the same conjugacy class are connected by equations ofthe <strong>for</strong>m B = X −1 AX. It follows that their matrix representations are connectedby corresponding equations of the <strong>for</strong>m D(B) =D(X −1 )D(A)D(X),<strong>and</strong>sobytheargument just given their representations will have equal traces <strong>and</strong> hence equalcharacters. Thus elements in the same conjugacy class have the same characters,1092


29.6 CHARACTERS3m I A, B C, D, EA 1 1 1 1 z; z 2 ; x 2 + y 2A 2 1 1 −1 R zE 2 −1 0 (x, y); (xz, yz); (R x ,R y ); (x 2 − y 2 , 2xy)Table 29.1 The character table <strong>for</strong> the irreps of group 3m (C 3v or S 3 ). Theright-h<strong>and</strong> column lists some common functions that trans<strong>for</strong>m according tothe irrep against which each is shown (see text).though, in general, these will vary from one representation to another. However,it might also happen that two or more conjugacy classes have the same charactersin a representation – indeed, in the trivial irrep A 1 , see (29.12), every elementinevitably has the character 1.For the irrep A 2 of the group 3m, the classes {I}, {A, B} <strong>and</strong> {C,D,E} havecharacters 1, 1 <strong>and</strong> −1, respectively, whilst they have characters 2, −1 <strong>and</strong>0respectively in irrep E.We are thus able to draw up a character table <strong>for</strong> the group 3m as shownin table 29.1. This table holds in compact <strong>for</strong>m most of the important in<strong>for</strong>mationon the behaviour of functions under the two-dimensional rotational <strong>and</strong>reflection symmetries of an equilateral triangle, i.e. under the elements of group3m. The entry under I <strong>for</strong> any irrep gives the dimension of the irrep, since itis equal to the trace of the unit matrix whose dimension is equal to that ofthe irrep. In other words, <strong>for</strong> the λth irrep χ (λ) (I) =n λ ,wheren λ is its dimension.In the extreme right-h<strong>and</strong> column we list some common functions of Cartesiancoordinates that trans<strong>for</strong>m, under the group 3m, according to the irrep on whoseline they are listed. Thus, as we have seen, z, z 2 ,<strong>and</strong>x 2 + y 2 are all unchangedby the group operations (though x <strong>and</strong> y individually are affected) <strong>and</strong> so arelisted against the one-dimensional irrep A 1 . Each of the pairs (x, y), (xz, yz), <strong>and</strong>(x 2 − y 2 , 2xy), however, is mixed as a pair by some of the operations, <strong>and</strong> so thesepairs are listed against the two-dimensional irrep E: each pair <strong>for</strong>ms a basis set<strong>for</strong> this irrep.The quantities R x , R y <strong>and</strong> R z refer to rotations about the indicated axes;they trans<strong>for</strong>m in the same way as the corresponding components of angularmomentum J, <strong>and</strong> their behaviour can be established by examining how thecomponents of J = r × p trans<strong>for</strong>m under the operations of the group. To dothis explicitly is beyond the scope of this book. However, it can be noted thatR z , being listed opposite the one-dimensional A 2 , is unchanged by I <strong>and</strong> by therotations A <strong>and</strong> B but changes sign under the mirror reflections C, D, <strong>and</strong>E, aswould be expected.1093


REPRESENTATION THEORY29.6.1 Orthogonality property of charactersSome of the most important properties of characters can be deduced from theorthogonality theorem (29.13),∑ [ ]ˆD (λ) ∗ [(X) ˆD (X)](µ) = g δ ik δ jl δ λµ .ijkl n λXIf we set j = i <strong>and</strong> l = k, so that both factors in any particular term in thesummation refer to diagonal elements of the representative matrices, <strong>and</strong> thensum both sides over i <strong>and</strong> k, we obtain∑ ∑n λ n∑ µ [ ]ˆD (λ) ∗ [(X) ˆD (µ)ii(X)]kk = g ∑n λn λXi=1 k=1n µ∑δ ik δ ik δ λµ .i=1 k=1Expressed in term of characters, this reads∑ [χ (λ) (X) ] ∗χ (µ) (X) = g ∑n λδniiδ 2 λµ = g ∑n λ1 × δ λµ = gδ λµ .λ n λXi=1i=1(29.14)In words, the (g-component) ‘vectors’ <strong>for</strong>med from the characters of the variousirreps of a group are mutually orthogonal, but each one has a squared magnitude(the sum of the squares of its components) equal to the order of the group.Since, as noted in the previous subsection, group elements in the same classhave the same characters, (29.14) can be written as a sum over classes rather thanelements. If c i denotes the number of elements in class C i <strong>and</strong> X i any element ofC i ,then∑ [c i χ (λ) (X i ) ] ∗χ (µ) (X i )=gδ λµ . (29.15)iAlthough we do not prove it here, there also exists a ‘completeness’ relation <strong>for</strong>characters. It makes a statement about the products of characters <strong>for</strong> a fixed pairof group elements, X 1 <strong>and</strong> X 2 , when the products are summed over all possibleirreps of the group. This is the converse of the summation process defined by(29.14). The completeness relation states that∑ [χ (λ) (X 1 ) ] ∗χ (λ) (X 2 )= g δ C1Cc2, (29.16)1λwhere element X 1 belongs to conjugacy class C 1 <strong>and</strong> X 2 belongs to C 2 . Thus thesum is zero unless X 1 <strong>and</strong> X 2 belong to the same class. For table 29.1 we canverify that these results are valid.(i) For ˆD (λ) = ˆD (µ) =A 1 or A 2 , (29.15) reads1(1) + 2(1) + 3(1) = 6,1094


29.7 COUNTING IRREPS USING CHARACTERSwhilst <strong>for</strong> ˆD (λ) = ˆD (µ) = E, it gives1(2 2 ) + 2(1) + 3(0) = 6.(ii) For ˆD (λ) =A 2 <strong>and</strong> ˆD (µ) = E, say, (29.15) reads1(1)(2) + 2(1)(−1) + 3(−1)(0) = 0.(iii) For X 1 = A <strong>and</strong> X 2 = D, say, (29.16) reads1(1) + 1(−1) + (−1)(0) = 0,whilst <strong>for</strong> X 1 = C <strong>and</strong> X 2 = E, both of which belong to class C 3 <strong>for</strong> whichc 3 =3,1(1) + (−1)(−1) + (0)(0) = 2 = 6 3 .29.7 Counting irreps using charactersThe expression of a general representation D = {D(X)} in terms of irreps, asgiven in (29.11), can be simplified by going from the full matrix <strong>for</strong>m to that ofcharacters. ThusD(X) =m 1 ˆD (1) (X) ⊕ m 2 ˆD (2) (X) ⊕ ···⊕ m N ˆD (N) (X)becomes, on taking the trace of both sides,χ(X) =N∑m λ χ (λ) (X). (29.17)λ=1Given the characters of the irreps of the group G to which the elements X belong,<strong>and</strong> the characters of the representation D = {D(X)}, theg equations (29.17)can be solved as simultaneous equations in the m λ , either by inspection or bymultiplying both sides by [ χ (µ) (X) ] ∗<strong>and</strong> summing over X, making use of (29.14)<strong>and</strong> (29.15), to obtainm µ = 1 ∑ [χ (µ) (X) ] ∗ 1 ∑ [χ(X) = c i χ (µ) (X i ) ] ∗χ(Xi ). (29.18)ggXiThat an unambiguous <strong>for</strong>mula can be given <strong>for</strong> each m λ , once the characterset (the set of characters of each of the group elements or, equivalently, ofeach of the conjugacy classes) of D is known, shows that, <strong>for</strong> any particulargroup, two representations with the same characters are equivalent. This stronglysuggests something that can be shown, namely, the number of irreps = the numberof conjugacy classes. The argument is as follows. Equation (29.17) is a set ofsimultaneous equations <strong>for</strong> N unknowns, the m λ , some of which may be zero. Thevalue of N is equal to the number of irreps of G. There are g different values ofX, but the number of different equations is only equal to the number of distinct1095


REPRESENTATION THEORYconjugacy classes, since any two elements of G in the same class have the samecharacter set <strong>and</strong> there<strong>for</strong>e generate the same equation. For a unique solutionto simultaneous equations in N unknowns, exactly N independent equations areneeded. Thus N is also the number of classes, establishing the stated result.◮Determine the irreps contained in the representation of the group 3m in the vector spacespanned by the functions x 2 , y 2 , xy.We first note that although these functions are not orthogonal they <strong>for</strong>m a basis set <strong>for</strong> arepresentation, since they are linearly independent quadratic <strong>for</strong>ms in x <strong>and</strong> y <strong>and</strong> any otherquadratic <strong>for</strong>m can be written (uniquely) in terms of them. We must establish how theytrans<strong>for</strong>m under the symmetry operations of group 3m. Weneedtodosoonly<strong>for</strong>arepresentativeelement of each conjugacy class, <strong>and</strong> naturally we take the simplest in each case.ThefirstclasscontainsonlyI (as always) <strong>and</strong> clearly D(I) isthe3× 3 unit matrix.The second class contains the rotations, A <strong>and</strong> B, <strong>and</strong> we choose to find D(A). Since,under A,x → − 1 √ √332 x + 2 y <strong>and</strong> y → − 2 x − 1 2 y,it follows thatx 2 → 1 4 x2 − √ 3xy + 32 4 y2 , y 2 → 3 4 x2 + √ 3xy + 12 4 y2 (29.19)<strong>and</strong>xy → √ 34 x2 − 1 xy − √ 32 4 y2 . (29.20)Hence D(A) can be deduced <strong>and</strong> is given below.The third <strong>and</strong> final class contains the reflections, C, D <strong>and</strong> E; oftheseC is much theeasiest to deal with. Under C, x →−x <strong>and</strong> y → y, causingxy to change sign but leavingx 2 <strong>and</strong> y 2 unaltered. The three matrices needed are thus⎛⎛D(I) =I 3 , D(C) = ⎝ 1 0 0⎞0 1 0 ⎠ , D(A) = ⎜⎝0 0 −114343− √ 34 2√1 34 2√34− √ 34− 1 2their traces are respectively 3, 1 <strong>and</strong> 0.It should be noticed that much more work has been done here than is necessary, sincethe traces can be computed immediately from the effects of the symmetry operations on thebasis functions. All that is needed is the weight of each basis function in the trans<strong>for</strong>medexpression <strong>for</strong> that function; these are clearly 1, 1, 1 <strong>for</strong> I, <strong>and</strong> 1 , 1 , − 1 <strong>for</strong> A, from (29.19)4 4 2<strong>and</strong> (29.20), <strong>and</strong> 1, 1, −1 <strong>for</strong>C, from the observations made just above the displayedmatrices. The traces are then the sums of these weights. The off-diagonal elements of thematrices need not be found, nor need the matrices be written out.From (29.17) we now need to find a superposition of the characters of the irreps thatgives representation D in the bottom line of table 29.2.By inspection it is obvious that D =A 1 ⊕ E, but we can use (29.18) <strong>for</strong>mally:m A1 = 1 [1(1)(3) + 2(1)(0) + 3(1)(1)] = 1,6m A2 = 1 [1(1)(3) + 2(1)(0) + 3(−1)(1)] = 0,6m E = 1 [1(2)(3) + 2(−1)(0) + 3(0)(1)] = 1.6Thus A 1 <strong>and</strong> E appear once each in the reduction of D, <strong>and</strong>A 2 not at all. Table 29.1gives the further in<strong>for</strong>mation, not needed here, that it is the combination x 2 + y 2 thattrans<strong>for</strong>ms as a one-dimensional irrep <strong>and</strong> the pair (x 2 − y 2 , 2xy) that<strong>for</strong>msabasisofthe two-dimensional irrep, E. ◭1096⎞⎟⎠ ;


29.7 COUNTING IRREPS USING CHARACTERSClassesIrrep I AB CDEA 1 1 1 1A 2 1 1 −1E 2 −1 0D 3 0 1Table 29.2 The characters of the irreps of the group 3m <strong>and</strong> of the representationD, which must be a superposition of some of them.29.7.1 Summation rules <strong>for</strong> irrepsThe first summation rule <strong>for</strong> irreps is a simple restatement of (29.14), with µ setequal to λ; itthenreads∑ [χ (λ) (X) ] ∗χ (λ) (X) =g.XIn words, the sum of the squares (modulus squared if necessary) of the charactersof an irrep taken over all elements of the group adds up to the order of thegroup. For group 3m (table 29.1), this takes the following explicit <strong>for</strong>ms:<strong>for</strong> A 1 , 1(1 2 )+2(1 2 )+3(1 2 )=6;<strong>for</strong> A 2 , 1(1 2 )+2(1 2 )+3(−1) 2 =6;<strong>for</strong> E, 1(2 2 )+2(−1) 2 +3(0 2 )=6.We next prove a theorem that is concerned not with a summation within an irrepbut with a summation over irreps.Theorem. If n µ is the dimension of the µth irrep of a group G then∑n 2 µ = g,where g is the order of the group.µProof. Define a representation of the group in the following way. Rearrangethe rows of the multiplication table of the group so that whilst the elements ina particular order head the columns, their inverses in the same order head therows. In this arrangement of the g × g table, the leading diagonal is entirelyoccupied by the identity element. Then, <strong>for</strong> each element X of the group, take asrepresentative matrix the multiplication-table array obtained by replacing X by1 <strong>and</strong> all other element symbols by 0. The matrices D reg (X) so obtained <strong>for</strong>m theregular representation of G; theyareeachg × g, have a single non-zero entry ‘1’in each row <strong>and</strong> column <strong>and</strong> (as will be verified by a little experimentation) have1097


REPRESENTATION THEORY(a)I A BI I A BA A B IB B I A(b)I A BI I A BB B I AA A B ITable 29.3 (a) The multiplication table of the cyclic group of order 3, <strong>and</strong>(b) its reordering used to generate the regular representation of the group.the same multiplication structure as the group G itself, i.e. they <strong>for</strong>m a faithfulrepresentation of G.Although not part of the proof, a simple example may help to make theseideas more transparent. Consider the cyclic group of order 3. Its multiplicationtable is shown in table 29.3(a) (a repeat of table 28.10(a) of the previous chapter),whilst table 29.3(b) shows the same table reordered so that the columns arestill labelled in the order I, A, B but the rows are now labelled in the orderI −1 = I, A −1 = B, B −1 = A. The three matrices of the regular representation arethen⎛D reg (I) = ⎝1 0⎞00 1 00 0 1⎛⎠ , D reg (A) = ⎝0 1⎞00 0 11 0 0⎛⎠ , D reg (B) = ⎝0 0 1⎞1 0 0 ⎠ .0 1 0An alternative, more mathematical, definition of the regular representation of agroup is[D reg (G k ) ] {1 ifij = Gk G j = G i ,0 otherwise.We now return to the proof. With the construction given, the regular representationhascharactersasfollows:χ reg (I) =g, χ reg (X) =0 ifX ≠ I.We now apply (29.18) to D reg to obtain <strong>for</strong> the number m µ of times that the irrepˆD (µ) appears in D reg (see 29.11))m µ = 1 ∑ [χ (µ) (X) ] ∗χ reg (X) = 1 [χ (µ) (I) ] ∗χ reg (I) = 1 ggg n µg = n µ .XThus an irrep ˆD (µ) of dimension n µ appears n µ times in D reg , <strong>and</strong> so by countingthe total number of basis functions, or by considering χ reg (I), we can conclude1098


29.7 COUNTING IRREPS USING CHARACTERSthat∑n 2 µ = g. (29.21)µThis completes the proof.As be<strong>for</strong>e, our st<strong>and</strong>ard demonstration group 3m provides an illustration. Inthis case we have seen already that there are two one-dimensional irreps <strong>and</strong> onetwo-dimensional irrep. This is in accord with (29.21) since1 2 +1 2 +2 2 =6, which is the order g of the group.Another straight<strong>for</strong>ward application of the relation (29.21), to the group withmultiplication table 29.3(a), yields immediate results. Since g =3,noneofitsirreps can have dimension 2 or more, as 2 2 = 4 is too large <strong>for</strong> (29.21) to besatisfied. Thus all irreps must be one-dimensional <strong>and</strong> there must be three ofthem (consistent with the fact that each element is in a class of its own, <strong>and</strong> thatthere are there<strong>for</strong>e three classes). The three irreps are the sets of 1 × 1 matrices(numbers)A 1 = {1, 1, 1} A 2 = {1,ω,ω 2 } A ∗ 2 = {1,ω 2 ,ω},where ω =exp(2πi/3); since the matrices are 1 × 1, the same set of nine numberswould be, of course, the entries in the character table <strong>for</strong> the irreps of the group.The fact that the numbers in each irrep are all cube roots of unity is discussedbelow. As will be noticed, two of these irreps are complex – an unusual occurrencein most applications – <strong>and</strong> <strong>for</strong>m a complex conjugate pair of one-dimensionalirreps. In practice, they function much as a two-dimensional irrep, but this is tobe ignored <strong>for</strong> <strong>for</strong>mal purposes such as theorems.A further property of characters can be derived from the fact that all elementsin a conjugacy class have the same order. Suppose that the element X has orderm, i.e.X m = I. This implies <strong>for</strong> a representation D of dimension n that[D(X)] m = I n . (29.22)Representations equivalent to D are generated as be<strong>for</strong>e by using similaritytrans<strong>for</strong>mations of the <strong>for</strong>mD Q (X) =Q −1 D(X)Q.In particular, if we choose the columns of Q to be the eigenvectors of D(X) then,as discussed in chapter 8,⎛⎞λ 1 0 ··· 0D Q (X) =0 λ. 2 ⎜ .⎝. ⎟. .. 0 ⎠0 ··· 0 λ n1099


REPRESENTATION THEORYwhere the λ i are the eigenvalues of D(X). There<strong>for</strong>e, from (29.22), we have that⎛λ m ⎞ ⎛⎞1 0 ··· 0 1 0 ··· 0..0 λ m 2 .⎜ .⎝. ⎟. .. 0 ⎠ = 0 1 .⎜ .⎝. ⎟. .. 0 ⎠ .0 ··· 0 λ m n 0 ··· 0 1Hence all the eigenvalues λ i are mth roots of unity, <strong>and</strong> so χ(X), the trace ofD(X), is the sum of n of these. In view of the implications of Lagrange’s theorem(section 28.6 <strong>and</strong> subsection 28.7.2), the only values of m allowed are the divisorsof the order g of the group.29.8 Construction of a character tableIn order to decompose representations into irreps on a routine basis usingcharacters, it is necessary to have available a character table <strong>for</strong> the group inquestion. Such a table gives, <strong>for</strong> each irrep µ of the group, the character χ (µ) (X)of the class to which group element X belongs. To construct such a table thefollowing properties of a group, established earlier in this chapter, may be used:(i) the number of classes equals the number of irreps;(ii) the ‘vector’ <strong>for</strong>med by the characters from a given irrep is orthogonal tothe ‘vector’ <strong>for</strong>med by the characters from a different irrep;(iii) ∑ µ n2 µ = g, wheren µ is the dimension of the µth irrep <strong>and</strong> g is the orderof the group;(iv) the identity irrep (one-dimensional with all characters equal to 1) is present<strong>for</strong> every group;(v) ∑ ∣X∣χ (µ) (X) ∣ 2 = g.(vi) χ (µ) (X) isthesumofn µ mth roots of unity, where m is the order of X.◮Construct the character table <strong>for</strong> the group 4mm (or C 4v ) using the properties of classes,irreps <strong>and</strong> characters so far established.The group 4mm is the group of two-dimensional symmetries of a square, namely rotationsof 0, π/2, π <strong>and</strong> 3π/2 <strong>and</strong> reflections in the mirror planes parallel to the coordinate axes<strong>and</strong> along the main diagonals. These are illustrated in figure 29.3. For this group there areeight elements:• the identity, I;• rotations by π/2 <strong>and</strong>3π/2, R <strong>and</strong> R ′ ;• arotationbyπ, Q ;• four mirror reflections m x , m y , m d <strong>and</strong> m d ′.Requirements (i) to (iv) at the start of this section put tight constraints on the possiblecharacter sets, as the following argument shows.The group is non-Abelian (clearly Rm x ≠ m x R), <strong>and</strong> so there are fewer than eightclasses, <strong>and</strong> hence fewer than eight irreps. But requirement (iii), with g = 8, then implies1100


29.8 CONSTRUCTION OF A CHARACTER TABLEm d ′m xm dm yFigure 29.3 The mirror planes associated with 4mm, the group of twodimensionalsymmetries of a square.that at least one irrep has dimension 2 or greater. However, there can be no irrep withdimension 3 or greater, since 3 2 > 8, nor can there be more than one two-dimensionalirrep, since 2 2 +2 2 = 8 would rule out a contribution to the sum in (iii) of 1 2 from theidentity irrep, <strong>and</strong> this must be present. Thus the only possibility is one two-dimensionalirrep <strong>and</strong>, to make the sum in (iii) correct, four one-dimensional irreps.There<strong>for</strong>e using (i) we can now deduce that there are five classes. This same conclusioncan be reached by evaluating X −1 YX <strong>for</strong> every pair of elements in G, as in the descriptionof conjugacy classes given in the previous chapter. However, it is tedious to do so <strong>and</strong>certainly much longer than the above. The five classes are I, Q, {R, R ′ }, {m x , m y }, {m d , m d ′}.It is straight<strong>for</strong>ward to show that only I <strong>and</strong> Q commute with every element of thegroup, so they are the only elements in classes of their own. Each other class must haveat least 2 members, but, as there are three classes to accommodate 8 − 2 = 6 elements,there must be exactly 2 in each class. This does not pair up the remaining 6 elements, butdoes say that the five classes have 1, 1, 2, 2, <strong>and</strong> 2 elements. Of course, if we had startedby dividing the group into classes, we would know the number of elements in each classdirectly.We cannot entirely ignore the group structure (though it sometimes happens that theresults are independent of the group structure – <strong>for</strong> example, all non-Abelian groups o<strong>for</strong>der 8 have the same character table!); thus we need to note in the present case thatm 2 i = I <strong>for</strong> i = x, y, d or d ′ <strong>and</strong>, as can be proved directly, Rm i = m i R ′ <strong>for</strong> the same fourvalues of label i. We also recall that <strong>for</strong> any pair of elements X <strong>and</strong> Y , D(XY )=D(X)D(Y ).We may conclude the following <strong>for</strong> the one-dimensional irreps.(a) In view of result (vi), χ(m i )=D(m i )=±1.(b) Since R 4 = I, result (vi) requires that χ(R) is one of 1, i, −1, −i. But, sinceD(R)D(m i )=D(m i )D(R ′ ), <strong>and</strong> the D(m i ) are just numbers, D(R) =D(R ′ ). FurtherD(R)D(R) =D(R)D(R ′ )=D(RR ′ )=D(I) =1,<strong>and</strong> so D(R) =±1 =D(R ′ ).(c) D(Q) =D(RR) =D(R)D(R) =1.If we add this to the fact that the characters of the identity irrep A 1 are all unity then wecan fill in those entries in character table 29.4 shown in bold.Suppose now that the three missing entries in a one-dimensional irrep are p, q <strong>and</strong> r,where each can only be ±1. Then, allowing <strong>for</strong> the numbers in each class, orthogonality1101


REPRESENTATION THEORY4mm I Q R, R ′ m x , m y m d , m d ′A 1 1 1 1 1 1A 2 1 1 1 −1 −1B 1 1 1 −1 1 −1B 2 1 1 −1 −1 1E 2 −2 0 0 0Table 29.4 The character table deduced <strong>for</strong> the group 4mm. For an explanationof the entries in bold see the text.with the characters of A 1 requires that1(1)(1) + 1(1)(1) + 2(1)(p) + 2(1)(q) + 2(1)(r) =0.The only possibility is that two of p, q, <strong>and</strong>r equal −1 <strong>and</strong> the other equals +1. Thiscan be achieved in three different ways, corresponding to the need to find three furtherdifferent one-dimensional irreps. Thus the first four lines of entries in character table 29.4can be completed. The final line can be completed by requiring it to be orthogonal to theother four. Property (v) has not been used here though it could have replaced part of theargument given. ◭29.9 Group nomenclatureThe nomenclature of published character tables, as we have said be<strong>for</strong>e, is erratic<strong>and</strong> sometimes un<strong>for</strong>tunate; <strong>for</strong> example, often E is used to represent, not onlya two-dimensional irrep, but also the identity operation, where we have used I.Thus the symbol E might appear in both the column <strong>and</strong> row headings of atable, though with quite different meanings in the two cases. In this book we useroman capitals to denote irreps.One-dimensional irreps are regularly denoted by A <strong>and</strong> B, B being used if arotation about the principal axis of 2π/n has character −1. Here n is the highestinteger such that a rotation of 2π/n is a symmetry operation of the system, <strong>and</strong>the principal axis is the one about which this occurs. For the group of operationson a square, n = 4, the axis is the perpendicular to the square <strong>and</strong> the rotationin question is R. The names <strong>for</strong> the group, 4mm <strong>and</strong> C 4v , derive from the factthat here n is equal to 4. Similarly, <strong>for</strong> the operations on an equilateral triangle,n = 3 <strong>and</strong> the group names are 3m <strong>and</strong> C 3v , but because the rotation by 2π/3 hascharacter +1 in all its one-dimensional irreps (see table 29.1), only A appears inthe irrep list.Two-dimensional irreps are denoted by E, as we have already noted, <strong>and</strong> threedimensionalirreps by T, although in many cases the symbols are modified byprimes <strong>and</strong> other alphabetic labels to denote variations in behaviour from oneirrep to another in respect of mirror reflections <strong>and</strong> parity inversions. In the studyof molecules, alternative names based on molecular angular momentum propertiesare common. It is beyond the scope of this book to list all these variations, or to1102


29.10 PRODUCT REPRESENTATIONSgive a large selection of character tables; our aim is to demonstrate <strong>and</strong> justifythe use of those found in the literature specifically dedicated to crystal physics ormolecular chemistry.Variations in notation are not restricted to the naming of groups <strong>and</strong> theirirreps, but extend to the symbols used to identify a typical element, <strong>and</strong> henceall members, of a conjugacy class in a group. In physics these are usually of thetypes n z , ¯n z or m x . The first of these denotes a rotation of 2π/n about the z-axis,<strong>and</strong> the second the same thing followed by parity inversion (all vectors r go to−r), whilst the third indicates a mirror reflection in a plane, in this case the planex =0.Typical chemistry symbols <strong>for</strong> classes are NC n , NC 2 n, NC x n , NS n , σ v , σ xy .Herethe first symbol N, where it appears, shows that there are N elements in theclass (a useful feature). The subscript n has the same meaning as in the physicsnotation, but σ rather than m is used <strong>for</strong> a mirror reflection, subscripts v, d or h orsuperscripts xy, xz or yz denoting the various orientations of the relevant mirrorplanes. Symmetries involving parity inversions are denoted by S; thus S n is thechemistry analogue of ¯n. None of what is said in this <strong>and</strong> the previous paragraphshould be taken as definitive, but merely as a warning of common variations innomenclature <strong>and</strong> as an initial guide to corresponding entities. Be<strong>for</strong>e using anyset of group character tables, the reader should ensure that he or she underst<strong>and</strong>sthe precise notation being employed.29.10 Product representationsIn quantum mechanical investigations we are often faced with the calculation ofwhat are called matrix elements. These normally take the <strong>for</strong>m of integrals over allspace of the product of two or more functions whose analytic <strong>for</strong>ms depend on themicroscopic properties (usually angular momentum <strong>and</strong> its components) of theelectrons or nuclei involved. For ‘bonding’ calculations involving ‘overlap integrals’there are usually two functions involved, whilst <strong>for</strong> transition probabilities a thirdfunction, giving the spatial variation of the interaction Hamiltonian, also appearsunder the integral sign.If the environment of the microscopic system under investigation has somesymmetry properties, then sometimes these can be used to establish, withoutdetailed evaluation, that the multiple integral must have zero value. We nowexpress the essential content of these ideas in group theoretical language.Suppose we are given an integral of the <strong>for</strong>m∫∫J = Ψφdτ or J = Ψξφdτto be evaluated over all space in a situation in which the physical system isinvariant under a particular group G of symmetry operations. For the integral to1103


REPRESENTATION THEORYbe non-zero the integr<strong>and</strong> must be invariant under each of these operations. Ingroup theoretical language, the integr<strong>and</strong> must trans<strong>for</strong>m as the identity, the onedimensionalrepresentation A 1 of G; more accurately, some non-vanishing part ofthe integr<strong>and</strong> must do so.An alternative way of saying this is that if under the symmetry operationsof G the integr<strong>and</strong> trans<strong>for</strong>ms according to a representation D <strong>and</strong> D does notcontain A 1 amongst its irreps then the integral J is necessarily zero. It should benoted that the converse is not true; J maybezeroevenifA 1 is present, since theintegral, whilst showing the required invariance, may still have the value zero.It is evident that we need to establish how to find the irreps that go to makeup a representation of a double or triple product when we already know theirreps according to which the factors in the product trans<strong>for</strong>m. The method isestablished by the following theorem.Theorem. For each element of a group the character in a product representation isthe product of the corresponding characters in the separate representations.Proof. Suppose that {u i } <strong>and</strong> {v j } are two sets of basis functions, that trans<strong>for</strong>munder the operations of a group G according to representations D (λ) <strong>and</strong> D (µ)respectively. Denote by u <strong>and</strong> v the corresponding basis vectors <strong>and</strong> let X be anelement of the group. Then the functions generated from u i <strong>and</strong> v j by the actionof X are calculated as follows, using (29.1) <strong>and</strong> (29.4):[ (Du ′ i = Xu i =(λ) (X) ) ]Tu = [ D (λ) (X) ]iii u i + ∑ [ (D (λ) (X) ) ] T u l,ill≠i[ (Dv j ′ = Xv j =(µ) (X) ) ]Tv = [ D (µ) (X) ]jjj v j + ∑ [ (D (µ) (X) ) ] T v m.jmm≠jHere [D(X)] ij is just a single element of the matrix D(X) <strong>and</strong>[D(X)] kk =[D T (X)] kkis simply a diagonal element from the matrix – the repeated subscript does notindicate summation. Now, if we take as basis functions <strong>for</strong> a product representationD prod (X) the products w k = u i v j (where the n λ n µ various possible pairs ofvalues i, j are labelled by k), we have also thatw k ′ = Xw k = Xu i v j =(Xu i )(Xv j )= [ D (λ) (X) ] [ii D (µ) (X) ] jj u iv j + terms not involving the product u i v j .This is to be compared with[ (Dw k ′ = Xw k =prod (X) ) ]Tw = [ D prod (X) ]kkk w k + ∑ [ (D prod (X) ) ] T w n,knn≠kwhere D prod (X) is the product representation matrix <strong>for</strong> element X of the group.The comparison shows that[D prod (X) ] kk = [ D (λ) (X) ] [ii D (µ) (X) ] jj .1104


29.11 PHYSICAL APPLICATIONS OF GROUP THEORYIt follows thatn λn µ∑χ prod [(X) = D prod (X) ] kkk=1n λ n µ∑ ∑ [= D (λ) (X) ] [ii D (µ) (X) ] jj=i=1 j=1{nλ∑ [D (λ) (X) ] iii=1}{ nµ∑ [D (µ) (X) ] }jjj=1= χ (λ) (X) χ (µ) (X). (29.23)This proves the theorem, <strong>and</strong> a similar argument leads to the corresponding result<strong>for</strong> integr<strong>and</strong>s in the <strong>for</strong>m of a product of three or more factors.An immediate corollary is that an integral whose integr<strong>and</strong> is the product oftwo functions trans<strong>for</strong>ming according to two different irreps is necessarily zero. Tosee this, we use (29.18) to determine whether irrep A 1 appears in the productcharacter set χ prod (X):m A1= 1 ∑ [χ(A 1) (X) ] ∗χ prod (X) = 1 ∑χ prod (X) = 1 ∑χ (λ) (X)χ (µ) (X).gg gXXXWe have used the fact that χ (A1) (X) =1<strong>for</strong>allX but now note that, by virtue of(29.14), the expression on the right of this equation is equal to zero unless λ = µ.Any complications due to non-real characters have been ignored – in practice,they are h<strong>and</strong>led automatically as it is usually Ψ ∗ φ, rather than Ψφ, that appearsin integr<strong>and</strong>s, though many functions are real in any case, <strong>and</strong> nearly all charactersare.Equation (29.23) is a general result <strong>for</strong> integr<strong>and</strong>s but, specifically in the contextof chemical bonding, it implies that <strong>for</strong> the possibility of bonding to exist, thetwo quantum wavefunctions must trans<strong>for</strong>m according to the same irrep. This isdiscussed further in the next section.29.11 Physical applications of group theoryAs we indicated at the start of chapter 28 <strong>and</strong> discussed in a little more detail atthe beginning of the present chapter, some physical systems possess symmetriesthat allow the results of the present chapter to be used in their analysis. Weconsider now some of the more common sorts of problem in which these resultsfind ready application.1105


REPRESENTATION THEORY1y42x3Figure 29.4manganese.A molecule consisting of four atoms of iodine <strong>and</strong> one of29.11.1 Bonding in moleculesWe have just seen that whether chemical bonding can take place in a moleculeis strongly dependent upon whether the wavefunctions of the two atoms <strong>for</strong>minga bond trans<strong>for</strong>m according to the same irrep. Thus it is sometimes useful to beable to find a wavefunction that does trans<strong>for</strong>m according to a particular irrepof a group of trans<strong>for</strong>mations. This can be done if the characters of the irrep areknown <strong>and</strong> a sensible starting point can be guessed. We state without proof thatstarting from any n-dimensional basis vector Ψ ≡ (Ψ 1 Ψ 2 ··· Ψ n ) T ,where{Ψ i }is a set of wavefunctions, the new vector Ψ (λ) ≡ (Ψ (λ)1Ψ (λ)2··· Ψ n (λ) ) T generatedbyΨ (λ)i= ∑ Xχ (λ)∗ (X)XΨ i (29.24)will trans<strong>for</strong>m according to the λth irrep. If the r<strong>and</strong>omly chosen Ψ happens notto contain any component that trans<strong>for</strong>ms in the desired way then the Ψ (λ) sogenerated is found to be a zero vector <strong>and</strong> it is necessary to select a new startingvector. An illustration of the use of this ‘projection operator’ is given in the nextexample.◮Consider a molecule made up of four iodine atoms lying at the corners of a square in thexy-plane, with a manganese atom at its centre, as shown in figure 29.4. Investigate whetherthe molecular orbital given by the superposition of p-state (angular momentum l = 1)atomic orbitalsΨ 1 =Ψ y (r − R 1 )+Ψ x (r − R 2 ) − Ψ y (r − R 3 ) − Ψ x (r − R 4 )can bond to the d-state atomic orbitals of the manganese atom described by either (i) φ 1 =(3z 2 −r 2 )f(r) or (ii) φ 2 =(x 2 −y 2 )f(r), wheref(r) is a function of r <strong>and</strong> so is unchanged byany of the symmetry operations of the molecule. Such linear combinations of atomic orbitalsare known as ring orbitals.We have eight basis functions, the atomic orbitals Ψ x (N) <strong>and</strong>Ψ y (N), where N =1, 2, 3, 4<strong>and</strong> indicates the position of an iodine atom. Since the wavefunctions are those of p-statesthey have the <strong>for</strong>ms xf(r) oryf(r) <strong>and</strong> lie in the directions of the x- <strong>and</strong>y-axes shown inthe figure. Since r is not changed by any of the symmetry operations, f(r) can be treated asa constant. The symmetry group of the system is 4mm, whose character table is table 29.4.1106


29.11 PHYSICAL APPLICATIONS OF GROUP THEORYCase (i). The manganese atomic orbital φ 1 =(3z 2 − r 2 )f(r), lying at the centre of themolecule, is not affected by any of the symmetry operations since z <strong>and</strong> r are unchangedby them. It clearly trans<strong>for</strong>ms according to the identity irrep A 1 . We there<strong>for</strong>e need toknow which combination of the iodine orbitals Ψ x (N) <strong>and</strong>Ψ y (N), if any, also trans<strong>for</strong>msaccording to A 1 .We use the projection operator (29.24). If we choose Ψ x (1) as the arbitrary onedimensionalstarting vector, we un<strong>for</strong>tunately obtain zero (as the reader may wish toverify), but Ψ y (1) is found to generate a new non-zero one-dimensional vector trans<strong>for</strong>mingaccording to A 1 . The results of acting on Ψ y (1) with the various symmetry elements Xcan be written down by inspection (see the discussion in section 29.2). So, <strong>for</strong> example, theΨ y (1) orbital centred on iodine atom 1 <strong>and</strong> aligned along the positive y-axisischangedby the anticlockwise rotation of π/2 produced by R ′ into an orbital centred on atom 4<strong>and</strong> aligned along the negative x-axis; thus R ′ Ψ y (1) = −Ψ x (4). The complete set of groupactions on Ψ y (1) is:I, Ψ y (1); Q, −Ψ y (3); R, Ψ x (2); R ′ , −Ψ x (4);m x , Ψ y (1); m y , −Ψ y (3); m d , Ψ x (2); m d ′, −Ψ x (4).Now χ (A1) (X) =1<strong>for</strong>allX, so (29.24) states that the sum of the above results <strong>for</strong> XΨ y (1),all with weight 1, gives a vector (here, since the irrep is one-dimensional, just a wavefunction)that trans<strong>for</strong>ms according to A 1 <strong>and</strong> is there<strong>for</strong>e capable of <strong>for</strong>ming a chemicalbond with the manganese wavefunction φ 1 .ItisΨ (A1) =2[Ψ y (1) − Ψ y (3) + Ψ x (2) − Ψ x (4)],though, of course, the factor 2 is irrelevant. This is precisely the ring orbital Ψ 1 given inthe problem, but here it is generated rather than guessed be<strong>for</strong>eh<strong>and</strong>.Case (ii). The atomic orbital φ 2 =(x 2 − y 2 )f(r) behaves as follows under the action oftypical conjugacy class members:I, φ 2 ; Q, φ 2 ; R, (y 2 − x 2 )f(r) =−φ 2 ; m x ,φ 2 ; m d , −φ 2 .From this we see that φ 2 trans<strong>for</strong>ms as a one-dimensional irrep, but, from table 29.4, thatirrep is B 1 not A 1 (the irrep according to which Ψ 1 trans<strong>for</strong>ms, as already shown). Thusφ 2 <strong>and</strong> Ψ 1 cannot <strong>for</strong>m a bond. ◭The original question did not ask <strong>for</strong> the the ring orbital to which φ 2 maybond, but it can be generated easily by using the values of XΨ y (1) calculated incase (i) <strong>and</strong> now weighting them according to the characters of B 1 :Ψ (B1) =Ψ y (1) − Ψ y (3) + (−1)Ψ x (2) − (−1)Ψ x (4)+Ψ y (1) − Ψ y (3) + (−1)Ψ x (2) − (−1)Ψ x (4)=2[Ψ y (1) − Ψ x (2) − Ψ y (3) + Ψ x (4)].Now we will find the other irreps of 4mm present in the space spanned bythe basis functions Ψ x (N) <strong>and</strong>Ψ y (N); at the same time this will illustrate theimportant point that since we are working with characters we are only interestedin the diagonal elements of the representative matrices. This means (section 29.2)that if we work in the natural representation D nat we need consider only thosefunctions that trans<strong>for</strong>m, wholly or partially, into themselves. Since we have noneed to write out the matrices explicitly, their size (8 × 8) is no drawback. All theirreps spanned by the basis functions Ψ x (N) <strong>and</strong>Ψ y (N) can be determined byconsidering the actions of the group elements upon them, as follows.1107


REPRESENTATION THEORY(i) Under I all eight basis functions are unchanged, <strong>and</strong> χ(I) =8.(ii) The rotations R, R ′ <strong>and</strong> Q change the value of N in every case <strong>and</strong> soall diagonal elements of the natural representation are zero <strong>and</strong> χ(R) =χ(Q) =0.(iii) m x takes x into −x <strong>and</strong> y into y <strong>and</strong>, <strong>for</strong> N = 1 <strong>and</strong> 3, leaves N unchanged,with the consequences (remember the <strong>for</strong>ms of Ψ x (N) <strong>and</strong>Ψ y (N)) thatΨ x (1) →−Ψ x (1), Ψ x (3) →−Ψ x (3),Ψ y (1) → Ψ y (1), Ψ y (3) → Ψ y (3).Thus χ(m x ) has four non-zero contributions, −1, −1, 1 <strong>and</strong> 1, togetherwith four zero contributions. The total is thus zero.(iv) m d <strong>and</strong> m d ′ leave no atom unchanged <strong>and</strong> so χ(m d )=0.The character set of the natural representation is thus 8, 0, 0, 0, 0, which, eitherby inspection or by applying <strong>for</strong>mula (29.18), shows thatD nat =A 1 ⊕ A 2 ⊕ B 1 ⊕ B 2 ⊕ 2E,i.e. that all possible irreps are present. We have constructed previously thecombinations of Ψ x (N) <strong>and</strong> Ψ y (N) that trans<strong>for</strong>m according to A 1 <strong>and</strong> B 1 .The others can be found in the same way.29.11.2 Matrix elements in quantum mechanicsIn section 29.10 we outlined the procedure <strong>for</strong> determining whether a matrixelement that involves the product of three factors as an integr<strong>and</strong> is necessarilyzero. We now illustrate this with a specific worked example.◮Determine whether a ‘dipole’ matrix element of the <strong>for</strong>m∫J = Ψ d1 xΨ d2 dτ,where Ψ d1 <strong>and</strong> Ψ d2 are d-state wavefunctions of the <strong>for</strong>ms xyf(r) <strong>and</strong> (x 2 − y 2 )g(r) respectively,can be non-zero (i) inamoleculewithsymmetryC 3v (or 3m), such as ammonia, <strong>and</strong>(ii) in a molecule with symmetry C 4v (or 4mm), such as the MnI 4 molecule considered inthe previous example.We will need to make reference to the character tables of the two groups. The table <strong>for</strong>C 3v is table 29.1 (section 29.6); that <strong>for</strong> C 4v is reproduced as table 29.5 from table 29.4 butwith the addition of another column showing how some common functions trans<strong>for</strong>m.We make use of (29.23), extended to the product of three functions. No attention needbe paid to f(r) <strong>and</strong>g(r) as they are unaffected by the group operations.Case (i). From the character table 29.1 <strong>for</strong> C 3v , we see that each of xy, x <strong>and</strong> x 2 − y 2<strong>for</strong>ms part of a basis set trans<strong>for</strong>ming according to the two-dimensional irrep E. Thus wemay fill in the array of characters (using chemical notation <strong>for</strong> the classes, except thatwe continue to use I rather than E) as shown in table 29.6. The last line is obtained by1108


29.11 PHYSICAL APPLICATIONS OF GROUP THEORY4mm I Q R, R ′ m x , m y m d , m d ′A 1 1 1 1 1 1 z; z 2 ; x 2 + y 2A 2 1 1 1 −1 −1 R zB 1 1 1 −1 1 −1 x 2 − y 2B 2 1 1 −1 −1 1 xyE 2 −2 0 0 0 (x, y); (xz, yz); (R x ,R y )Table 29.5 The character table <strong>for</strong> the irreps of group 4mm (or C 4v ). Theright-h<strong>and</strong> column lists some common functions, or, <strong>for</strong> the two-dimensionalirrep E, pairs of functions, that trans<strong>for</strong>m according to the irrep against whichthey are shown.Function Irrep ClassesI 2C 3 3σ vxy E 2 −1 0x E 2 −1 0x 2 − y 2 E 2 −1 0product 8 −1 0Table 29.6 The character sets, <strong>for</strong> the group C 3v (or 3mm), of three functions<strong>and</strong> of their product x 2 y(x 2 − y 2 ).Function Irrep ClassesI C 2 2C 6 2σ v 2σ dxy B 2 1 1 −1 −1 1x E 2 −2 0 0 0x 2 − y 2 B 1 1 1 −1 1 −1product 2 −2 0 0 0Table 29.7 The character sets, <strong>for</strong> the group C 4v (or 4mm), of three functions,<strong>and</strong> of their product x 2 y(x 2 − y 2 ).multiplying together the corresponding characters <strong>for</strong> each of the three elements. Now, byinspection, or by applying (29.18), i.e.m A1 = 1 [1(1)(8) + 2(1)(−1) + 3(1)(0)] = 1,6we see that irrep A 1 does appear in the reduced representation of the product, <strong>and</strong> so Jis not necessarily zero.Case (ii). From table 29.5 we find that, under the group C 4v , xy <strong>and</strong> x 2 − y 2 trans<strong>for</strong>mas irreps B 2 <strong>and</strong> B 1 respectively <strong>and</strong> that x is part of a basis set trans<strong>for</strong>ming as E. Thusthe calculation table takes the <strong>for</strong>m of table 29.7 (again, chemical notation <strong>for</strong> the classeshas been used).Here inspection is sufficient, as the product is exactly that of irrep E <strong>and</strong> irrep A 1 iscertainly not present. Thus J is necessarily zero <strong>and</strong> the dipole matrix element vanishes. ◭1109


REPRESENTATION THEORYy 3x 3y 1y2x 1x 2Figure 29.5An equilateral array of masses <strong>and</strong> springs.29.11.3 Degeneracy of normal modesAs our final area <strong>for</strong> illustrating the usefulness of group theoretical results weconsider the normal modes of a vibrating system (see chapter 9). This analysishas far-reaching applications in physics, chemistry <strong>and</strong> engineering. For a givensystem, normal modes that are related by some symmetry operation have the samefrequency of vibration; the modes are said to be degenerate. It can be shown thatsuch modes span a vector space that trans<strong>for</strong>ms according to some irrep of thegroup G of symmetry operations of the system. Moreover, the degeneracy ofthe modes equals the dimension of the irrep. As an illustration, we consider thefollowing example.◮Investigate the possible vibrational modes of the equilateral triangular arrangement ofequal masses <strong>and</strong> springs shown in figure 29.5. Demonstrate that two are degenerate.Clearly the symmetry group is that of the symmetry operations on an equilateral triangle,namely 3m (or C 3v ), whose character table is table 29.1. As on a previous occasion, it ismost convenient to use the natural representation D nat of this group (it almost alwayssaves having to write out matrices explicitly) acting on the six-dimensional vector space(x 1 , y 1 , x 2 , y 2 , x 3 , y 3 ). In this example the natural <strong>and</strong> regular representations coincide, butthis is not usually the case.We note that in table 29.1 the second class contains the rotations A (by π/3) <strong>and</strong> B (by2π/3), also known as R <strong>and</strong> R ′ . This class is known as 3 z in crystallographic notation, orC 3 in chemical notation, as explained in section 29.9. The third class contains C, D, E, thethree mirror reflections.Clearly χ(I) = 6. Since all position labels are changed by a rotation, χ(3 z )=0.Forthemirror reflections the simplest representative class member to choose is the reflection m y inthe plane containing the y 3 -axis, since then only label 3 is unchanged; under m y , x 3 →−x 3<strong>and</strong> y 3 → y 3 , leading to the conclusion that χ(m y ) = 0. Thus the character set is 6, 0, 0.Using (29.18) <strong>and</strong> the character table 29.1 shows thatD nat =A 1 ⊕ A 2 ⊕ 2E.1110


29.11 PHYSICAL APPLICATIONS OF GROUP THEORYHowever, we have so far allowed x i , y i to be completely general, <strong>and</strong> we must now identify<strong>and</strong> remove those irreps that do not correspond to vibrations. These will be the irrepscorresponding to bodily translations of the triangle <strong>and</strong> to its rotation without relativemotion of the three masses.Bodily translations are linear motions of the centre of mass, which has coordinatesx =(x 1 + x 2 + x 3 )/3 <strong>and</strong> y =(y 1 + y 2 + y 3 )/3).Table 29.1 shows that such a coordinate pair (x, y) trans<strong>for</strong>ms according to the twodimensionalirrep E; this accounts <strong>for</strong> one of the two such irreps found in the naturalrepresentation.It can be shown that, as stated in table 29.1, planar bodily rotations of the triangle –rotations about the z-axis, denoted by R z – trans<strong>for</strong>m as irrep A 2 . Thus, when the linearmotions of the centre of mass, <strong>and</strong> pure rotation about it, are removed from our reducedrepresentation, we are left with E⊕A 1 . So, E <strong>and</strong> A 1 must be the irreps corresponding to theinternal vibrations of the triangle – one doubly degenerate mode <strong>and</strong> one non-degeneratemode.The physical interpretation of this is that two of the normal modes of the system havethe same frequency <strong>and</strong> one normal mode has a different frequency (barring accidentalcoincidences <strong>for</strong> other reasons). It may be noted that in quantum mechanics the energyquantum of a normal mode is proportional to its frequency. ◭In general, group theory does not tell us what the frequencies are, since it isentirely concerned with the symmetry of the system <strong>and</strong> not with the values ofmasses <strong>and</strong> spring constants. However, using this type of reasoning, the resultsfrom representation theory can be used to predict the degeneracies of atomicenergy levels <strong>and</strong>, given a perturbation whose Hamiltonian (energy operator) hassome degree of symmetry, the extent to which the perturbation will resolve thedegeneracy. Some of these ideas are explored a little further in the next section<strong>and</strong> in the exercises.29.11.4 Breaking of degeneraciesIf a physical system has a high degree of symmetry, invariant under a group G ofreflections <strong>and</strong> rotations, say, then, as implied above, it will normally be the casethat some of its eigenvalues (of energy, frequency, angular momentum etc.) aredegenerate. However, if a perturbation that is invariant only under the operationsof the elements of a smaller symmetry group (a subgroup of G) is added, some ofthe original degeneracies may be broken. The results derived from representationtheory can be used to decide the extent of the degeneracy-breaking.The normal procedure is to use an N-dimensional basis vector, consisting of theN degenerate eigenfunctions, to generate an N-dimensional representation of thesymmetry group of the perturbation. This representation is then decomposed intoirreps. In general, eigenfunctions that trans<strong>for</strong>m according to different irreps nolonger share the same frequency of vibration. We illustrate this with the followingexample.1111


REPRESENTATION THEORYMMMFigure 29.6masses.A circular drumskin loaded with three symmetrically placed◮A circular drumskin has three equal masses placed on it at the vertices of an equilateraltriangle, as shown in figure 29.6. Determine which degenerate normal modes of the drumskincan be split in frequency by this perturbation.When no masses are present the normal modes of the drum-skin are either non-degenerateor two-fold degenerate (see chapter 21). The degenerate eigenfunctions Ψ of the nth normalmode have the <strong>for</strong>msJ n (kr)(cos nθ)e ±iωt or J n (kr)(sin nθ)e ±iωt .There<strong>for</strong>e, as explained above, we need to consider the two-dimensional vector spacespanned by Ψ 1 =sinnθ <strong>and</strong> Ψ 2 =cosnθ. This will generate a two-dimensional representationof the group 3m (or C 3v ), the symmetry group of the perturbation. Taking the easiestelement from each of the three classes (identity, rotations, <strong>and</strong> reflections) of group 3m,we haveIΨ 1 =Ψ 1 , IΨ 2 =Ψ 2 ,AΨ 1 =sin [ n ( θ − 2 π)] = ( cos 2 nπ) Ψ3 3 1 − ( sin 2 nπ) Ψ3 2 ,AΨ 2 =cos [ n ( θ − 2 π)] = ( cos 2 nπ) Ψ3 3 2 + ( sin 2 nπ) Ψ3 1 ,CΨ 1 = sin[n(π − θ)] = −(cos nπ)Ψ 1 ,CΨ 2 =cos[n(π − θ)] = (cos nπ)Ψ 2 .The three representative matrices are there<strong>for</strong>e(cos23D(I) =I 2 , D(A) =nπ − sin 2 nπ)( )3 − cos nπ 0sin 2 nπ cos 2 nπ , D(C) =.0 cosnπ3 3The characters of this representation are χ(I) =2,χ(A) =2cos(2nπ/3) <strong>and</strong> χ(C) =0.Using (29.18) <strong>and</strong> table 29.1, we find that(m A1 = 1 6 2+4cos2nπ) = m3 A2(m E = 1 6 4 − 4cos2nπ) .3Thus{A 1 ⊕ A 2 if n =3, 6, 9, ...,D =E otherwise.Hence the normal modes n =3, 6, 9, ... each trans<strong>for</strong>m under the operations of 3m1112


29.12 EXERCISESas the sum of two one-dimensional irreps <strong>and</strong>, using the reasoning given in the previousexample, are there<strong>for</strong>e split in frequency by the perturbation. For other values of n therepresentation is irreducible <strong>and</strong> so the degeneracy cannot be split. ◭29.12 Exercises29.1 A group G hasfourelementsI,X,Y <strong>and</strong> Z, which satisfy X 2 = Y 2 = Z 2 =XY Z = I. Show that G is Abelian <strong>and</strong> hence deduce the <strong>for</strong>m of its charactertable.Show that the matrices(1 0D(I) =0 1D(Y )=(−1 −p0 1)(−1 0, D(X) =0 −1), D(Z) =(1 p0 −1where p is a real number, <strong>for</strong>m a representation D of G. Find its characters <strong>and</strong>decompose it into irreps.29.2 Using a square whose corners lie at coordinates (±1, ±1), <strong>for</strong>m a natural representationof the dihedral group D 4 . Find the characters of the representation,<strong>and</strong>, using the in<strong>for</strong>mation (<strong>and</strong> class order) in table 29.4 (p. 1102), express therepresentation in terms of irreps.Now <strong>for</strong>m a representation in terms of eight 2 × 2 orthogonal matrices, byconsidering the effect of each of the elements of D 4 on a general vector (x, y).Confirm that this representation is one of the irreps found using the naturalrepresentation.29.3 The quaternion group Q (see exercise 28.20) has eight elements {±1, ±i, ±j,±k}obeying the relationsi 2 = j 2 = k 2 = −1, ij = k = −ji.Determine the conjugacy classes of Q <strong>and</strong> deduce the dimensions of its irreps.Show that Q is homomorphic to the four-element group V, which is generated bytwo distinct elements a <strong>and</strong> b with a 2 = b 2 =(ab) 2 = I. Find the one-dimensionalirreps of V <strong>and</strong> use these to help determine the full character table <strong>for</strong> Q.29.4 Construct the character table <strong>for</strong> the irreps of the permutation group S 4 asfollows.(a) By considering the possible <strong>for</strong>ms of its cycle notation, determine the numberof elements in each conjugacy class of the permutation group S 4 ,<strong>and</strong>showthat S 4 has five irreps. Give the logical reasoning that shows they must consistof two three-dimensional, one two-dimensional, <strong>and</strong> two one-dimensionalirreps.(b) By considering the odd <strong>and</strong> even permutations in the group S 4 , establish thecharacters <strong>for</strong> one of the one-dimensional irreps.(c) Form a natural matrix representation of 4 × 4 matrices based on a set ofobjects {a, b, c, d}, which may or may not be equal to each other, <strong>and</strong>, byselecting one example from each conjugacy class, show that this naturalrepresentation has characters 4, 2, 1, 0, 0. In the four-dimensional vectorspace in which each of the four coordinates takes on one of the four valuesa, b, c or d, the one-dimensional subspace consisting of the four points withcoordinates of the <strong>for</strong>m {a, a, a, a} is invariant under the permutation group<strong>and</strong> hence trans<strong>for</strong>ms according to the invariant irrep A 1 . The remainingthree-dimensional subspace is irreducible; use this <strong>and</strong> the characters deducedabove to establish the characters <strong>for</strong> one of the three-dimensional irreps, T 1 .1113),),


REPRESENTATION THEORY(d) Complete the character table using orthogonality properties, <strong>and</strong> check thesummation rule <strong>for</strong> each irrep. You should obtain table 29.8.Typical element <strong>and</strong> class sizeIrrep (1) (12) (123) (1234) (12)(34)1 6 8 6 3A 1 1 1 1 1 1A 2 1 −1 1 −1 1E 2 0 −1 0 2T 1 3 1 0 −1 −1T 2 3 −1 0 1 −1Table 29.8 The character table <strong>for</strong> the permutation group S 4 .29.5 In exercise 28.10, the group of pure rotations taking a cube into itself was foundto have 24 elements. The group is isomorphic to the permutation group S 4 ,considered in the previous question, <strong>and</strong> hence has the same character table, oncecorresponding classes have been established. By counting the number of elementsin each class, make the correspondences below (the final two cannot be decidedpurely by counting, <strong>and</strong> should be taken as given).Permutation Symbol Actionclass type (physics)(1) I none(123) 3 rotations about a body diagonal(12)(34) 2 z rotation of π about the normal to a face(1234) 4 z rotations of ±π/2 about the normal to a face(12) 2 d rotation of π about an axis through thecentres of opposite edgesRe<strong>for</strong>mulate the character table 29.8 in terms of the elements of the rotationsymmetry group (432 or O) of a cube <strong>and</strong> use it when answering exercises 29.7<strong>and</strong> 29.8.29.6 Consider a regular hexagon orientated so that two of its vertices lie on the x-axis.Find matrix representations of a rotation R through 2π/6 <strong>and</strong> a reflection m y inthe y-axis by determining their effects on vectors lying in the xy-plane . Showthat a reflection m x in the x-axis can be written as m x = m y R 3 , <strong>and</strong> that the 12elements of the symmetry group of the hexagon are given by R n or R n m y .Using the representations of R <strong>and</strong> m y as generators, find a two-dimensionalrepresentation of the symmetry group, C 6 , of the regular hexagon. Is it a faithfulrepresentation?29.7 In a certain crystalline compound, a thorium atom lies at the centre of a regularoctahedron of six sulphur atoms at positions (±a, 0, 0), (0, ±a, 0), (0, 0, ±a). Thesecan be considered as being positioned at the centres of the faces of a cube ofside 2a. The sulphur atoms produce at the site of the thorium atom an electricfield that has the same symmetry group as a cube (432 or O).The five degenerate d-electron orbitals of the thorium atom can be expressed,relative to any arbitrary polar axis, as(3 cos 2 θ − 1)f(r), e ±iφ sin θ cos θf(r), e ±2iφ sin 2 θf(r).A rotation about that polar axis by an angle φ ′ effectively changes φ to φ − φ ′ .1114


29.12 EXERCISESUse this to show that the character of the rotation in a representation based onthe orbital wavefunctions is given by1+2cosφ ′ +2cos2φ ′<strong>and</strong> hence that the characters of the representation, in the order of the symbolsgiven in exercise 29.5, is 5, −1, 1, −1, 1. Deduce that the five-fold degeneratelevel is split into two levels, a doublet <strong>and</strong> a triplet.29.8 Sulphur hexafluoride is a molecule with the same structure as the crystallinecompound in exercise 29.7, except that a sulphur atom is now the central atom.The following are the <strong>for</strong>ms of some of the electronic orbitals of the sulphuratom, together with the irreps according to which they trans<strong>for</strong>m under thesymmetry group 432 (or O).Ψ s = f(r) A 1Ψ p1 = zf(r) T 1Ψ d1 =(3z 2 − r 2 )f(r) EΨ d2 =(x 2 − y 2 )f(r) EΨ d3 = xyf(r) T 2The function x trans<strong>for</strong>ms according to the irrep T 1 . Use the above data todetermine whether dipole matrix elements of the <strong>for</strong>m J = ∫ φ 1 xφ 2 dτ can benon-zero <strong>for</strong> the following pairs of orbitals φ 1 ,φ 2 in a sulphur hexafluoridemolecule: (a) Ψ d1 , Ψ s ;(b)Ψ d1 , Ψ p1 ;(c)Ψ d2 , Ψ d1 ;(d)Ψ s , Ψ d3 ;(e)Ψ p1 , Ψ s .29.9 The hydrogen atoms in a methane molecule CH 4 <strong>for</strong>m a perfect tetrahedronwith the carbon atom at its centre. The molecule is most conveniently describedmathematically by placing the hydrogen atoms at the points (1, 1, 1), (1, −1, −1),(−1, 1, −1) <strong>and</strong> (−1, −1, 1). The symmetry group to which it belongs, the tetrahedralgroup (¯43m or T d ),hasclassestypifiedbyI, 3,2 z , m d <strong>and</strong> ¯4 z , where the firstthree are as in exercise 29.5, m d is a reflection in the mirror plane x − y =0<strong>and</strong>¯4 z is a rotation of π/2 about the z-axis followed by an inversion in the origin. Areflection in a mirror plane can be considered as a rotation of π about an axisperpendicular to the plane, followed by an inversion in the origin.Thecharactertable<strong>for</strong>thegroup¯43m is very similar to that <strong>for</strong> the group432, <strong>and</strong> has the <strong>for</strong>m shown in table 29.9.Typical element <strong>and</strong> class size Functions trans<strong>for</strong>mingIrreps I 3 2 z¯4z m d according to irrep1 8 3 6 6A 1 1 1 1 1 1 x 2 + y 2 + z 2A 2 1 1 1 −1 −1E 2 −1 2 0 0 (x 2 − y 2 , 3z 2 − r 2 )T 1 3 0 −1 1 −1 (R x ,R y ,R z )T 2 3 0 −1 −1 1 (x, y, z); (xy, yz, zx)Table 29.9The character table <strong>for</strong> group ¯43m.By following the steps given below, determine how many different internal vibrationfrequencies the CH 4 molecule has.(a) Consider a representation based on the twelve coordinates x i ,y i ,z i <strong>for</strong>i =1, 2, 3, 4. For those hydrogen atoms that trans<strong>for</strong>m into themselves, arotation through an angle θ about an axis parallel to one of the coordinateaxes gives rise in the natural representation to the diagonal elements 1 <strong>for</strong>1115


REPRESENTATION THEORYthe corresponding coordinate <strong>and</strong> 2 cos θ <strong>for</strong> the two orthogonal coordinates.If the rotation is followed by an inversion then these entries are multipliedby −1. Atoms not trans<strong>for</strong>ming into themselves give a zero diagonal contribution.Show that the characters of the natural representation are 12, 0, 0,0, 2 <strong>and</strong> hence that its expression in terms of irreps isA 1 ⊕ E ⊕ T 1 ⊕ 2T 2 .(b) The irreps of the bodily translational <strong>and</strong> rotational motions are included inthis expression <strong>and</strong> need to be identified <strong>and</strong> removed. Show that when thisis done it can be concluded that there are three different internal vibrationfrequencies in the CH 4 molecule. State their degeneracies <strong>and</strong> check thatthey are consistent with the expected number of normal coordinates neededto describe the internal motions of the molecule.29.10 Investigate the properties of an alternating group <strong>and</strong> construct its charactertable as follows.(a) The set of even permutations of four objects (a proper subgroup of S 4 )is known as the alternating group A 4 . List its twelve members using cyclenotation.(b) Assume that all permutations with the same cycle structure belong to thesame conjugacy class. Show that this leads to a contradiction, <strong>and</strong> hencedemonstrates that, even if two permutations have the same cycle structure,they do not necessarily belong to the same class.(c) By evaluating the productsp 1 = (123)(4) • (12)(34) • (132)(4) <strong>and</strong> p 2 = (132)(4) • (12)(34) • (123)(4)deduce that the three elements of A 4 with structure of the <strong>for</strong>m (12)(34)belong to the same class.(d) By evaluating products of the <strong>for</strong>m (1α)(βγ) • (123)(4) • (1α)(βγ), where α, β, γare various combinations of 2, 3, 4, show that the class to which (123)(4)belongs contains at least four members. Show the same <strong>for</strong> (124)(3).(e) By combining results (b), (c) <strong>and</strong> (d) deduce that A 4 has exactly four classes,<strong>and</strong> determine the dimensions of its irreps.(f) Using the orthogonality properties of characters <strong>and</strong> noting that elements ofthe <strong>for</strong>m (124)(3) have order 3, find the character table <strong>for</strong> A 4 .29.11 Use the results of exercise 28.23 to find the character table <strong>for</strong> the dihedral groupD 5 , the symmetry group of a regular pentagon.29.12 Demonstrate that equation (29.24) does, indeed, generate a set of vectors trans<strong>for</strong>mingaccording to an irrep λ, by sketching <strong>and</strong> superposing drawings of anequilateral triangle of springs <strong>and</strong> masses, based on that shown in figure 29.5.CCCA B AB 30 ◦ AB30 ◦ (a) (b) (c)Figure 29.7 The three normal vibration modes of the equilateral array. Mode(a) is known as the ‘breathing mode’. Modes (b) <strong>and</strong> (c) trans<strong>for</strong>m accordingto irrep E <strong>and</strong> have equal vibrational frequencies.1116


29.13 HINTS AND ANSWERS(a) Make an initial sketch showing an arbitrary small mass displacement from,say, vertex C. Draw the results of operating on this initial sketch with eachof the symmetry elements of the group 3m (C 3v ).(b) Superimpose the results, weighting them according to the characters of irrepA 1 (table 29.1 in section 29.6) <strong>and</strong> verify that the resultant is a symmetricalarrangement in which all three masses move symmetrically towards (or awayfrom) the centroid of the triangle. The mode is illustrated in figure 29.7(a).(c)Start again, this time considering a displacement δ of C parallel to the x-axis.Form a similar superposition of sketches weighted according to the charactersof irrep E (note that the reflections are not needed). The resultant containssome bodily displacement of the triangle, since this also trans<strong>for</strong>ms accordingto E. Show that the displacement of the centre of mass is ¯x = δ, ȳ =0.Subtract this out, <strong>and</strong> verify that the remainder is of the <strong>for</strong>m shown infigure 29.7(c).(d) Using an initial displacement parallel to the y-axis, <strong>and</strong> an analogous procedure,generate the remaining normal mode, degenerate with that in (c) <strong>and</strong>shown in figure 29.7(b).29.13 Further investigation of the crystalline compound considered in exercise 29.7shows that the octahedron is not quite perfect but is elongated along the (1, 1, 1)direction with the sulphur atoms at positions ±(a+δ,δ,δ), ±(δ,a+δ,δ), ±(δ,δ,a+δ), where δ ≪ a. This structure is invariant under the (crystallographic) symmetrygroup 32 with three two-fold axes along directions typified by (1, −1, 0). Thelatter axes, which are perpendicular to the (1, 1, 1) direction, are axes of twofoldsymmetry <strong>for</strong> the perfect octahedron. The group 32 is really the threedimensionalversion of the group 3m <strong>and</strong> has the same character table as table 29.1(section 29.6). Use this to show that, when the distortion of the octahedron isincluded, the doublet found in exercise 29.7 is unsplit but the triplet breaks upinto a singlet <strong>and</strong> a doublet.29.13 Hints <strong>and</strong> answers29.1 There are four classes <strong>and</strong> hence four one-dimensional irreps, which must haveentriesasfollows:1,1,1,1; 1,1,−1, −1; 1, −1, 1, −1; 1, −1, −1, 1. Thecharacters of D are 2, −2, 0, 0 <strong>and</strong> so the irreps present are the last two of these.29.3 There are five classes {1}, {−1}, {±i}, {±j}, {±k}; there are four one-dimensionalirreps <strong>and</strong> one two-dimensional irrep. Show that ab = ba. The homomorphismis ±1 → I, ±i → a, ±j → b, ±k → ab. V is Abelian <strong>and</strong> hence has fourone-dimensional irreps.In the class order given above, the characters <strong>for</strong> Q are as follows:ˆD (1) , 1, 1, 1, 1, 1; ˆD (2) , 1, 1, 1, −1, −1; ˆD (3) , 1, 1, −1, 1, −1;ˆD (4) , 1, 1, −1, −1, 1; ˆD (5) , 2, −2, 0, 0, 0.29.5 Note that the fourth <strong>and</strong> fifth classes each have 6 members.29.7 The five basis functions of the representation are multiplied by 1, e −iφ′ , e +iφ′ ,e −2iφ′ , e +2iφ′ as a result of the rotation. The character is the sum of these <strong>for</strong>rotations of 0, 2π/3, π, π/2, π; D rep =E+T 2 .29.9 (b) The bodily translation has irrep T 2 <strong>and</strong> the rotation has irrep T 1 . The irrepsof the internal vibrations are A 1 ,E,T 2 , with respective degeneracies 1, 2, 3,making six internal coordinates (12 in total, minus three translational, minusthree rotational).29.11 There are four classes <strong>and</strong> hence four irreps, which can only be the identityirrep, one other one-dimensional irrep, <strong>and</strong> two two-dimensional irreps. In theclass order {I}, {R,R 4 }, {R 2 ,R 3 }, {m i } the second one-dimensional irrep must1117


REPRESENTATION THEORY(because of orthogonality) have characters 1, 1, 1, −1. The summation rules<strong>and</strong> orthogonality require the other two character sets to be 2, (−1 + √ 5)/2,(−1 − √ 5)/2, 0 <strong>and</strong> 2, (−1 − √ 5)/2, (−1+ √ 5)/2, 0. Note that R has order 5 <strong>and</strong>that, e.g., (−1+ √ 5)/2 =exp(2πi/5) + exp(8πi/5).29.13 The doublet irrep E (characters 2, −1, 0) appears in both 432 <strong>and</strong> 32 <strong>and</strong> sois unsplit. The triplet T 1 (characters 3, 0, 1) splits under 32 into doublet E(characters 2, −1, 0) <strong>and</strong> singlet A 1 (characters 1, 1, 1).1118


30ProbabilityAll scientists will know the importance of experiment <strong>and</strong> observation <strong>and</strong>,equally, be aware that the results of some experiments depend to a degree onchance. For example, in an experiment to measure the heights of a r<strong>and</strong>om sampleof people, we would not be in the least surprised if all the heights were found tobe different; but, if the experiment were repeated often enough, we would expectto find some sort of regularity in the results. Statistics, which is the subject of thenext chapter, is concerned with the analysis of real experimental data of this sort.First, however, we discuss probability. To a pure mathematician, probability is anentirely theoretical subject based on axioms. Although this axiomatic approach isimportant, <strong>and</strong> we discuss it briefly, an approach to probability more in keepingwith its eventual applications in statistics is adopted here.We first discuss the terminology required, with particular reference to theconvenient graphical representation of experimental results as Venn diagrams.The concepts of r<strong>and</strong>om variables <strong>and</strong> distributions of r<strong>and</strong>om variables are thenintroduced. It is here that the connection with statistics is made; we assert thatthe results of many experiments are r<strong>and</strong>om variables <strong>and</strong> that those results havesome sort of regularity, which is represented by a distribution. Precise definitionsof a r<strong>and</strong>om variable <strong>and</strong> a distribution are then given, as are the definingequations <strong>for</strong> some important distributions. We also derive some useful quantitiesassociated with these distributions.30.1 Venn diagramsWe call a single per<strong>for</strong>mance of an experiment a trial <strong>and</strong> each possible resultan outcome. Thesample space S of the experiment is then the set of all possibleoutcomes of an individual trial. For example, if we throw a six-sided die then thereare six possible outcomes that together <strong>for</strong>m the sample space of the experiment.At this stage we are not concerned with how likely a particular outcome might1119


PROBABILITYAiiiiiiBSivFigure 30.1A Venn diagram.be (we will return to the probability of an outcome in due course) but ratherwill concentrate on the classification of possible outcomes. It is clear that somesample spaces are finite (e.g. the outcomes of throwing a die) whilst others areinfinite (e.g. the outcomes of measuring people’s heights). Most often, one is notinterested in individual outcomes but in whether an outcome belongs to a givensubset A (say) of the sample space S; these subsets are called events. For example,we might be interested in whether a person is taller or shorter than 180 cm, inwhich case we divide the sample space into just two events: namely, that theoutcome (height measured) is (i) greater than 180 cm or (ii) less than 180 cm.A common graphical representation of the outcomes of an experiment is theVenn diagram. A Venn diagram usually consists of a rectangle, the interior ofwhich represents the sample space, together with one or more closed curves insideit. The interior of each closed curve then represents an event. Figure 30.1 showsa typical Venn diagram representing a sample space S <strong>and</strong> two events A <strong>and</strong>B. Every possible outcome is assigned to an appropriate region; in this examplethere are four regions to consider (marked i to iv in figure 30.1):(i) outcomes that belong to event A but not to event B;(ii) outcomes that belong to event B but not to event A;(iii) outcomes that belong to both event A <strong>and</strong> event B;(iv) outcomes that belong to neither event A nor event B.◮A six-sided die is thrown. Let event A be ‘the number obtained is divisible by 2’ <strong>and</strong> eventB be ‘the number obtained is divisible by 3’. Draw a Venn diagram to represent these events.It is clear that the outcomes 2, 4, 6 belong to event A <strong>and</strong> that the outcomes 3, 6 belongto event B. Of these, 6 belongs to both A <strong>and</strong> B. The remaining outcomes, 1, 5, belong toneither A nor B. The appropriate Venn diagram is shown in figure 30.2. ◭In the above example, one outcome, 6, is divisible by both 2 <strong>and</strong> 3 <strong>and</strong> sobelongs to both A <strong>and</strong> B. This outcome is placed in region iii of figure 30.1, whichis called the intersection of A <strong>and</strong> B <strong>and</strong> is denoted by A ∩ B (see figure 30.3(a)).If no events lie in the region of intersection then A <strong>and</strong> B are said to be mutuallyexclusive or disjoint. In this case, often the Venn diagram is drawn so that theclosed curves representing the events A <strong>and</strong> B do not overlap, so as to make1120


30.1 VENN DIAGRAMSA42 6 3BS15Figure 30.2 The Venn diagram <strong>for</strong> the outcomes of the die-throwing trialsdescribed in the worked example.ABABS(a)S(b)ABAĀS(c)S(d)Figure 30.3 Venn diagrams: the shaded regions show (a) A ∩ B, theintersectionof two events A <strong>and</strong> B, (b)A ∪ B, the union of events A <strong>and</strong> B, (c)the complement Ā of an event A, (d)A − B, those outcomes in A that do notbelong to B.graphically explicit the fact that A <strong>and</strong> B are disjoint. It is not necessary, however,to draw the diagram in this way, since we may simply assign zero outcomes tothe shaded region in figure 30.3(a). An event that contains no outcomes is calledthe empty event <strong>and</strong> denoted by ∅. The event comprising all the elements thatbelong to either A or B, or to both, is called the union of A <strong>and</strong> B <strong>and</strong> is denotedby A ∪ B (see figure 30.3(b)). In the previous example, A ∪ B = {2, 3, 4, 6}.It is sometimes convenient to talk about those outcomes that do not belong toa particular event. The set of outcomes that do not belong to A is called thecomplement of A <strong>and</strong> is denoted by Ā (see figure 30.3(c)); this can also be writtenas Ā = S − A. It is clear that A ∪ Ā = S <strong>and</strong> A ∩ Ā = ∅.The above notation can be extended in an obvious way, so that A − B denotesthe outcomes in A that do not belong to B. It is clear from figure 30.3(d) thatA − B canalsobewrittenasA ∩ ¯B. Finally, when all the outcomes in event B(say) also belong to event A, but A may contain, in addition, outcomes that do1121


PROBABILITYA1B24756 3C8SFigure 30.4regions.The general Venn diagram <strong>for</strong> three events is divided into eightnot belong to B, thenB is called a subset of A, a situation that is denoted byB ⊂ A; alternatively, one may write A ⊃ B, which states that A contains B. Inthiscase, the closed curve representing the event B is often drawn lying completelywithin the closed curve representing the event A.The operations ∪ <strong>and</strong> ∩ are extended straight<strong>for</strong>wardly to more than twoevents. If there exist n events A 1 ,A 2 ,...,A n , in some sample space S, then theevent consisting of all those outcomes that belong to one or more of the A i is theunion of A 1 ,A 2 ,...,A n <strong>and</strong> is denoted byA 1 ∪ A 2 ∪ ···∪ A n . (30.1)Similarly, the event consisting of all the outcomes that belong to every one of theA i is called the intersection of A 1 ,A 2 ,...,A n <strong>and</strong> is denoted byIf, <strong>for</strong> any pair of values i, j with i ≠ j,A 1 ∩ A 2 ∩ ···∩ A n . (30.2)A i ∩ A j = ∅ (30.3)then the events A i <strong>and</strong> A j are said to be mutually exclusive or disjoint.Consider three events A, B <strong>and</strong> C with a Venn diagram such as is shown infigure 30.4. It will be clear that, in general, the diagram will be divided into eightregions <strong>and</strong> they will be of four different types. Three regions correspond to asingle event; three regions are each the intersection of exactly two events; oneregion is the three-fold intersection of all three events; <strong>and</strong> finally one regioncorresponds to none of the events. Let us now consider the numbers of differentregions in a general n-event Venn diagram.For one-event Venn diagrams there are two regions, <strong>for</strong> the two-event casethere are four regions <strong>and</strong>, as we have just seen, <strong>for</strong> the three-event case there areeight. In the general n-event case there are 2 n regions, as is clear from the factthat any particular region R lies either inside or outside the closed curve of anyparticular event. With two choices (inside or outside) <strong>for</strong> each of n closed curves,there are 2 n different possible combinations with which to characterise R. Oncen1122


30.1 VENN DIAGRAMSgets beyond three it becomes impossible to draw a simple two-dimensional Venndiagram, but this does not change the results.The 2 n regions will break down into n + 1 types, with the numbers of each typeas follows § no events, n C 0 =1;one event but no intersections, n C 1 = n;two-fold intersections, n C 2 = 1 2n(n − 1);three-fold intersections, n C 3 = 1 3!n(n − 1)(n − 2);.an n-fold intersection, n C n =1.That this makes a total of 2 n can be checked by considering the binomialexpansion2 n =(1+1) n =1+n + 1 2n(n − 1) + ···+1.Using Venn diagrams, it is straight<strong>for</strong>ward to show that the operations ∩ <strong>and</strong>∪ obey the following algebraic laws:commutativity, A ∩ B = B ∩ A, A ∪ B = B ∪ A;associativity, (A ∩ B) ∩ C = A ∩ (B ∩ C), (A ∪ B) ∪ C = A ∪ (B ∪ C);distributivity, A ∩ (B ∪ C) =(A ∩ B) ∪ (A ∩ C),A ∪ (B ∩ C) =(A ∪ B) ∩ (A ∪ C);idempotency, A ∩ A = A, A ∪ A = A.◮Show that (i) A ∪ (A ∩ B) =A ∩ (A ∪ B) =A, (ii) (A − B) ∪ (A ∩ B) =A.(i) Using the distributivity <strong>and</strong> idempotency laws above, we see thatA ∪ (A ∩ B) =(A ∪ A) ∩ (A ∪ B) =A ∩ (A ∪ B).By sketching a Venn diagram it is immediately clear that both expressions are equal toA. Nevertheless, we here proceed in a more <strong>for</strong>mal manner in order to deduce this resultalgebraically. Let us begin by writingX = A ∪ (A ∩ B) =A ∩ (A ∪ B), (30.4)from which we want to deduce a simpler expression <strong>for</strong> the event X. Using the first equalityin (30.4) <strong>and</strong> the algebraic laws <strong>for</strong> ∩ <strong>and</strong> ∪, wemaywriteA ∩ X = A ∩ [A ∪ (A ∩ B)]=(A ∩ A) ∪ [A ∩ (A ∩ B)]= A ∪ (A ∩ B) =X.§ The symbols n C i ,<strong>for</strong>i =0, 1, 2,...,n, are a convenient notation <strong>for</strong> combinations; they <strong>and</strong> theirproperties are discussed in chapter 1.1123


PROBABILITYSince A ∩ X = X we must have X ⊂ A. Now, using the second equality in (30.4) in asimilar way, we findA ∪ X = A ∪ [A ∩ (A ∪ B)]=(A ∪ A) ∩ [A ∪ (A ∪ B)]= A ∩ (A ∪ B) =X,from which we deduce that A ⊂ X. Thus,sinceX ⊂ A <strong>and</strong> A ⊂ X, we must conclude thatX = A.(ii) Since we do not know how to deal with compound expressions containing a minussign, we begin by writing A − B = A ∩ ¯B as mentioned above. Then, using the distributivitylaw, we obtain(A − B) ∪ (A ∩ B) =(A ∩ ¯B) ∪ (A ∩ B)= A ∩ ( ¯B ∪ B)= A ∩ S = A.In fact, this result, like the first one, can be proved trivially by drawing a Venn diagram. ◭Further useful results may be derived from Venn diagrams. In particular, it issimple to show that the following rules hold:(i) if A ⊂ B then Ā ⊃ ¯B;(ii) A ∪ B = Ā ∩ ¯B;(iii) A ∩ B = Ā ∪ ¯B.Statements (ii) <strong>and</strong> (iii) are known jointly as de Morgan’s laws <strong>and</strong> are sometimesuseful in simplifying logical expressions.◮There exist two events A <strong>and</strong> B such that(X ∪ A) ∪ (X ∪ Ā) =B.Find an expression <strong>for</strong> the event X in terms of A <strong>and</strong> B.We begin by taking the complement of both sides of the above expression: applying deMorgan’s laws we obtain¯B =(X ∪ A) ∩ (X ∪ Ā).We may then use the algebraic laws obeyed by ∩ <strong>and</strong> ∪ to yield¯B = X ∪ (A ∩ Ā) =X ∪ ∅ = X.Thus, we find that X = ¯B. ◭30.2 ProbabilityIn the previous section we discussed Venn diagrams, which are graphical representationsof the possible outcomes of experiments. We did not, however, giveany indication of how likely each outcome or event might be when any particularexperiment is per<strong>for</strong>med. Most experiments show some regularity. By this wemean that the relative frequency of an event is approximately the same on eachoccasion that a set of trials is per<strong>for</strong>med. For example, if we throw a die N1124


30.2 PROBABILITYtimes then we expect that a six will occur approximately N/6 times (assuming,of course, that the die is not biased). The regularity of outcomes allows us todefine the probability, Pr(A), as the expected relative frequency of event A in alarge number of trials. More quantitatively, if an experiment has a total of n Soutcomes in the sample space S, <strong>and</strong>n A of these outcomes correspond to theevent A, then the probability that event A will occur isPr(A) = n A. (30.5)n S30.2.1 Axioms <strong>and</strong> theoremsFrom (30.5) we may deduce the following properties of the probability Pr(A).(i) For any event A in a sample space S,0 ≤ Pr(A) ≤ 1. (30.6)If Pr(A) =1thenA is a certainty; if Pr(A) =0thenA is an impossibility.(ii) For the entire sample space S we havePr(S) = n S=1, (30.7)n Swhich simply states that we are certain to obtain one of the possibleoutcomes.(iii) If A <strong>and</strong> B are two events in S then, from the Venn diagrams in figure 30.3,we see thatn A∪B = n A + n B − n A∩B , (30.8)the final subtraction arising because the outcomes in the intersection ofA <strong>and</strong> B are counted twice when the outcomes of A are added to thoseof B. Dividing both sides of (30.8) by n S , we obtain the addition rule <strong>for</strong>probabilitiesPr(A ∪ B) =Pr(A)+Pr(B) − Pr(A ∩ B). (30.9)However, if A <strong>and</strong> B are mutually exclusive events (A ∩ B = ∅) thenPr(A ∩ B) = 0 <strong>and</strong> we obtain the special casePr(A ∪ B) =Pr(A)+Pr(B). (30.10)(iv) If Ā is the complement of A then Ā <strong>and</strong> A are mutually exclusive events.Thus, from (30.7) <strong>and</strong> (30.10) we have1=Pr(S) =Pr(A ∪ Ā) =Pr(A)+Pr(Ā),from which we obtain the complement lawPr(Ā) =1− Pr(A). (30.11)1125


PROBABILITYThis is particularly useful <strong>for</strong> problems in which evaluating the probabilityof the complement is easier than evaluating the probability of the eventitself.◮Calculate the probability of drawing an ace or a spade from a pack of cards.Let A be the event that an ace is drawn <strong>and</strong> B the event that a spade is drawn. Itimmediately follows that Pr(A) = 4 = 113<strong>and</strong> Pr(B) = = 1 . The intersection of A <strong>and</strong>52 13 52 4B consists of only the ace of spades <strong>and</strong> so Pr(A ∩ B) = 1 . Thus, from (30.9)52Pr(A ∪ B) = 1 + 1 − 1 = 4 .13 4 52 13In this case it is just as simple to recognise that there are 16 cards in the pack that satisfythe required condition (13 spades plus three other aces) <strong>and</strong> so the probability is 16 . ◭ 52The above theorems can easily be extended to a greater number of events. Forexample, if A 1 ,A 2 ,...,A n are mutually exclusive events then (30.10) becomesPr(A 1 ∪ A 2 ∪ ···∪ A n )=Pr(A 1 )+Pr(A 2 )+···+Pr(A n ). (30.12)Furthermore, if A 1 ,A 2 ,...,A n (whether mutually exclusive or not) exhaust S, i.e.are such that A 1 ∪ A 2 ∪ ···∪ A n = S, thenPr(A 1 ∪ A 2 ∪ ···∪ A n )=Pr(S) =1. (30.13)◮A biased six-sided die has probabilities 1 p, p, p, p, p, 2p of showing 1, 2, 3, 4, 5, 62respectively. Calculate p.Given that the individual events are mutually exclusive, (30.12) can be applied to givePr(1 ∪ 2 ∪ 3 ∪ 4 ∪ 5 ∪ 6) = 1 13p + p + p + p + p +2p = p.2 2The union of all possible outcomes on the LHS of this equation is clearly the samplespace, S, <strong>and</strong>soPr(S) = 13 p. 2Now using (30.7),132p =Pr(S) =1 ⇒ p = . ◭2 13When the possible outcomes of a trial correspond to more than two events,<strong>and</strong> those events are not mutually exclusive, the calculation of the probability ofthe union of a number of events is more complicated, <strong>and</strong> the generalisation ofthe addition law (30.9) requires further work. Let us begin by considering theunion of three events A 1 , A 2 <strong>and</strong> A 3 , which need not be mutually exclusive. Wefirst define the event B = A 2 ∪ A 3 <strong>and</strong>, using the addition law (30.9), we obtainPr(A 1 ∪ A 2 ∪ A 3 )=Pr(A 1 ∪ B) =Pr(A 1 )+Pr(B) − Pr(A 1 ∩ B).(30.14)1126


30.2 PROBABILITYHowever, we may write Pr(A 1 ∩ B) asPr(A 1 ∩ B) =Pr[A 1 ∩ (A 2 ∪ A 3 )]=Pr[(A 1 ∩ A 2 ) ∪ (A 1 ∩ A 3 )]=Pr(A 1 ∩ A 2 )+Pr(A 1 ∩ A 3 ) − Pr(A 1 ∩ A 2 ∩ A 3 ).Substituting this expression, <strong>and</strong> that <strong>for</strong> Pr(B) obtained from (30.9), into (30.14)we obtain the probability addition law <strong>for</strong> three general events,Pr(A 1 ∪ A 2 ∪ A 3 )=Pr(A 1 )+Pr(A 2 )+Pr(A 3 ) − Pr(A 2 ∩ A 3 ) − Pr(A 1 ∩ A 3 )− Pr(A 1 ∩ A 2 )+Pr(A 1 ∩ A 2 ∩ A 3 ). (30.15)◮Calculate the probability of drawing from a pack of cards one that is an ace or is a spadeor shows an even number (2,4,6,8,10).If, as previously, A is the event that an ace is drawn, Pr(A) = 4 . Similarly the event B,52that a spade is drawn, has Pr(B) = 13 . The further possibility C, that the card is even (but52not a picture card) has Pr(C) = 20 . The two-fold intersections have probabilities52Pr(A ∩ B) = 1 52 , Pr(A ∩ C) =0, Pr(B ∩ C) = 5 52 .There is no three-fold intersection as events A <strong>and</strong> C are mutually exclusive. HencePr(A ∪ B ∪ C) = 1 31[(4+13+20)− (1+0+5)+(0)] =52 52 .The reader should identify the 31 cards involved. ◭When the probabilities are combined to calculate the probability <strong>for</strong> the unionof the n general events, the result, which may be proved by induction upon n (seethe answer to exercise 30.4), isPr(A 1 ∪ A 2 ∪ ···∪ A n )= ∑ iPr(A i ) − ∑ i,jPr(A i ∩ A j )+ ∑ i,j,kPr(A i ∩ A j ∩ A k )− ···+(−1) n+1 Pr(A 1 ∩ A 2 ∩ ···∩ A n ). (30.16)Each summation runs over all possible sets of subscripts, except those in whichany two subscripts in a set are the same. The number of terms in the summationof probabilities of m-fold intersections of the n events is given by n C m (as discussedin section 30.1). Equation (30.9) is a special case of (30.16) in which n = 2 <strong>and</strong>only the first two terms on the RHS survive. We now illustrate this result with aworked example that has n = 4 <strong>and</strong> includes a four-fold intersection.1127


PROBABILITY◮Find the probability of drawing from a pack a card that has at least one of the followingproperties:A, itisanace;B, it is a spade;C, it is a black honour card (ace, king, queen, jack or 10);D, it is a black ace.Measuring all probabilities in units of 1 , the single-event probabilities are52Pr(A) =4, Pr(B) =13, Pr(C) =10, Pr(D) =2.The two-fold intersection probabilities, measured in the same units, arePr(A ∩ B) =1, Pr(A ∩ C) =2, Pr(A ∩ D) =2,Pr(B ∩ C) =5, Pr(B ∩ D) =1, Pr(C ∩ D) =2.The three-fold intersections have probabilitiesPr(A ∩ B ∩ C) =1, Pr(A ∩ B ∩ D) =1, Pr(A ∩ C ∩ D) =2, Pr(B ∩ C ∩ D) =1.Finally, the four-fold intersection, requiring all four conditions to hold, is satisfied only bythe ace of spades, <strong>and</strong> hence (again in units of 1 52Pr(A ∩ B ∩ C ∩ D) =1.Substituting in (30.16) givesP = 1 20[(4+13+10+2)− (1+2+2+5+1+2)+(1+1+2+1)− (1)] =52 52 . ◭We conclude this section on basic theorems by deriving a useful generalexpression <strong>for</strong> the probability Pr(A ∩ B) thattwoeventsA <strong>and</strong> B both occur inthe case where A (say) is the union of a set of n mutually exclusive events A i .Inthis caseA ∩ B =(A 1 ∩ B) ∪ ···∪ (A n ∩ B),where the events A i ∩ B are also mutually exclusive. Thus, from the addition law(30.12) <strong>for</strong> mutually exclusive events, we findPr(A ∩ B) = ∑ Pr(A i ∩ B). (30.17)iMoreover, in the special case where the events A i exhaust the sample space S, wehave A ∩ B = S ∩ B = B, <strong>and</strong> we obtain the total probability lawPr(B) = ∑ iPr(A i ∩ B). (30.18)30.2.2 Conditional probabilitySo far we have defined only probabilities of the <strong>for</strong>m ‘what is the probability thatevent A happens?’. In this section we turn to conditional probability, the probabilitythat a particular event occurs given the occurrence of another, possibly related,event. For example, we may wish to know the probability of event B, drawingan1128


30.2 PROBABILITYace from a pack of cards from which one has already been removed, given thatevent A, the card already removed was itself an ace, has occurred.We denote this probability by Pr(B|A) <strong>and</strong> may obtain a <strong>for</strong>mula <strong>for</strong> it byconsidering the total probability Pr(A ∩ B) =Pr(B ∩ A) that both A <strong>and</strong> B willoccur. This may be written in two ways, i.e.From this we obtainPr(A ∩ B) =Pr(A)Pr(B|A)=Pr(B)Pr(A|B).Pr(A|B) =Pr(A ∩ B)Pr(B)(30.19)<strong>and</strong>Pr(B ∩ A)Pr(B|A) = . (30.20)Pr(A)In terms of Venn diagrams, we may think of Pr(B|A) as the probability of B inthe reduced sample space defined by A. Thus, if two events A <strong>and</strong> B are mutuallyexclusive thenPr(A|B) =0=Pr(B|A). (30.21)When an experiment consists of drawing objects at r<strong>and</strong>om from a given setof objects, it is termed sampling a population. We need to distinguish betweentwo different ways in which such a sampling experiment may be per<strong>for</strong>med. Afteran object has been drawn at r<strong>and</strong>om from the set it may either be put asideor returned to the set be<strong>for</strong>e the next object is r<strong>and</strong>omly drawn. The <strong>for</strong>mer istermed ‘sampling without replacement’, the latter ‘sampling with replacement’.◮Find the probability of drawing two aces at r<strong>and</strong>om from a pack of cards (i) when thefirst card drawn is replaced at r<strong>and</strong>om into the pack be<strong>for</strong>e the second card is drawn, <strong>and</strong>(ii) when the first card is put aside after being drawn.Let A be the event that the first card is an ace, <strong>and</strong> B the event that the second card is anace. NowPr(A ∩ B) =Pr(A)Pr(B|A),<strong>and</strong> <strong>for</strong> both (i) <strong>and</strong> (ii) we know that Pr(A) = 4 = 1 . 52 13(i) If the first card is replaced in the pack be<strong>for</strong>e the next is drawn then Pr(B|A) =Pr(B) = 4 = 1 ,sinceA <strong>and</strong> B are independent events. We then have52 13Pr(A ∩ B) =Pr(A)Pr(B) = 1 13 × 113 = 1169 .(ii) If the first card is put aside <strong>and</strong> the second then drawn, A <strong>and</strong> B are not independent<strong>and</strong> Pr(B|A) = 3 , with the result that51Pr(A ∩ B) =Pr(A)Pr(B|A) = 1 13 × 351 = 1221 . ◭1129


PROBABILITYTwo events A <strong>and</strong> B are statistically independent if Pr(A|B) =Pr(A) (or equivalentlyif Pr(B|A) =Pr(B)). In words, the probability of A given B is then the sameas the probability of A regardless of whether B occurs. For example, if we throwa coin <strong>and</strong> a die at the same time, we would normally expect that the probabilityof throwing a six was independent of whether a head was thrown. If A <strong>and</strong> B arestatistically independent then it follows thatPr(A ∩ B) =Pr(A)Pr(B). (30.22)In fact, on the basis of intuition <strong>and</strong> experience, (30.22) may be regarded as thedefinition of the statistical independence of two events.The idea of statistical independence is easily extended to an arbitrary numberof events A 1 ,A 2 ,...,A n . The events are said to be (mutually) independent ifPr(A i ∩ A j )=Pr(A i )Pr(A j ),Pr(A i ∩ A j ∩ A k )=Pr(A i )Pr(A j )Pr(A k ),..Pr(A 1 ∩ A 2 ∩ ···∩ A n )=Pr(A 1 )Pr(A 2 ) ···Pr(A n ),<strong>for</strong> all combinations of indices i, j <strong>and</strong> k <strong>for</strong> which no two indices are the same.Even if all n events are not mutually independent, any two events <strong>for</strong> whichPr(A i ∩ A j )=Pr(A i )Pr(A j ) are said to be pairwise independent.We now derive two results that often prove useful when working with conditionalprobabilities. Let us suppose that an event A is the union of n mutuallyexclusive events A i .IfB is some other event then from (30.17) we havePr(A ∩ B) = ∑ Pr(A i ∩ B).iDividing both sides of this equation by Pr(B), <strong>and</strong> using (30.19), we obtainPr(A|B) = ∑ iPr(A i |B), (30.23)which is the addition law <strong>for</strong> conditional probabilities.Furthermore, if the set of mutually exclusive events A i exhausts the samplespace S then, from the total probability law (30.18), the probability Pr(B) of someevent B in S can be written asPr(B) = ∑ iPr(A i )Pr(B|A i ). (30.24)◮A collection of traffic isl<strong>and</strong>s connected by a system of one-way roads is shown in figure30.5. At any given isl<strong>and</strong> a car driver chooses a direction at r<strong>and</strong>om from those available.What is the probability that a driver starting at O will arrive at B?In order to leave O the driver must pass through one of A 1 , A 2 , A 3 or A 4 , which thus<strong>for</strong>m a complete set of mutually exclusive events. Since at each isl<strong>and</strong> (including O) thedriver chooses a direction at r<strong>and</strong>om from those available, we have that Pr(A i )= 1 <strong>for</strong> 41130


30.2 PROBABILITYA 4A 3OA 1A 2BFigure 30.5A collection of traffic isl<strong>and</strong>s connected by one-way roads.i =1, 2, 3, 4. From figure 30.5, we see also thatPr(B|A 1 )= 1 , Pr(B|A 3 2)= 1 , Pr(B|A 3 3)=0, Pr(B|A 4 )= 2 = 1 . 4 2Thus, using the total probability law (30.24), we find that the probability of arriving at Bis given byPr(B) = ∑ (Pr(A i )Pr(B|A i )= 1 1 + 1 +0+ ) 14 3 3 2 =7. ◭ 24iFinally, we note that the concept of conditional probability may be straight<strong>for</strong>wardlyextended to several compound events. For example, in the case of threeevents A, B, C, we may write Pr(A ∩ B ∩ C) in several ways, e.g.Pr(A ∩ B ∩ C) =Pr(C)Pr(A ∩ B|C)=Pr(B ∩ C)Pr(A|B ∩ C)=Pr(C)Pr(B|C)Pr(A|B ∩ C).◮Suppose {A i } is a set of mutually exclusive events that exhausts the sample space S. IfB<strong>and</strong> C are two other events in S, show thatPr(B|C) = ∑ iPr(A i |C)Pr(B|A i ∩ C).Using (30.19) <strong>and</strong> (30.17), we may writePr(C)Pr(B|C) =Pr(B ∩ C) = ∑ iPr(A i ∩ B ∩ C). (30.25)Each term in the sum on the RHS can be exp<strong>and</strong>ed as an appropriate product ofconditional probabilities,Pr(A i ∩ B ∩ C) =Pr(C)Pr(A i |C)Pr(B|A i ∩ C).Substituting this <strong>for</strong>m into (30.25) <strong>and</strong> dividing through by Pr(C) gives the requiredresult. ◭1131


PROBABILITY30.2.3 Bayes’ theoremIn the previous section we saw that the probability that both an event A <strong>and</strong> arelated event B will occur can be written either as Pr(A)Pr(B|A) orPr(B)Pr(A|B).Hencefrom which we obtain Bayes’ theorem,Pr(A)Pr(B|A) =Pr(B)Pr(A|B),Pr(A|B) = Pr(A) Pr(B|A). (30.26)Pr(B)This theorem clearly shows that Pr(B|A) ≠Pr(A|B), unless Pr(A) =Pr(B). It issometimes useful to rewrite Pr(B), if it is not known directly, asso that Bayes’ theorem becomesPr(B) =Pr(A)Pr(B|A)+Pr(Ā)Pr(B|Ā)Pr(A|B) =Pr(A)Pr(B|A)Pr(A)Pr(B|A) +Pr(Ā)Pr(B|Ā) . (30.27)◮Suppose that the blood test <strong>for</strong> some disease is reliable in the following sense: <strong>for</strong> peoplewho are infected with the disease the test produces a positive result in 99.99% of cases; <strong>for</strong>people not infected a positive test result is obtained in only 0.02% of cases. Furthermore,assume that in the general population one person in 10 000 people is infected. A person isselected at r<strong>and</strong>om <strong>and</strong> found to test positive <strong>for</strong> the disease. What is the probability thatthe individual is actually infected?Let A be the event that the individual is infected <strong>and</strong> B be the event that the individualtests positive <strong>for</strong> the disease. Using Bayes’ theorem the probability that a person who testspositive is actually infected isPr(A|B) =Pr(A)Pr(B|A)Pr(A)Pr(B|A) +Pr(Ā)Pr(B|Ā) .Now Pr(A) =1/10000 = 1 − Pr(Ā), <strong>and</strong> we are told that Pr(B|A) = 9999/10000 <strong>and</strong>Pr(B|Ā) =2/10000. Thus we obtain1/10000 × 9999/10000Pr(A|B) =(1/10000 × 9999/10000) + (9999/10000 × 2/10000) = 1 3 .Thus, there is only a one in three chance that a person chosen at r<strong>and</strong>om, who testspositive <strong>for</strong> the disease, is actually infected.At a first glance, this answer may seem a little surprising, but the reason <strong>for</strong> the counterintuitiveresult is that the probability that a r<strong>and</strong>omly selected person is not infected is9999/10000, which is very high. Thus, the 0.02% chance of a positive test <strong>for</strong> an uninfectedperson becomes significant. ◭1132


30.3 PERMUTATIONS AND COMBINATIONSWe note that (30.27) may be written in a more general <strong>for</strong>m if S is not simplydivided into A <strong>and</strong> Ā but, rather, into any set of mutually exclusive events A i thatexhaust S. Using the total probability law (30.24), we may then writePr(B) = ∑ iPr(A i )Pr(B|A i ),so that Bayes’ theorem takes the <strong>for</strong>mPr(A|B) =Pr(A)Pr(B|A)∑i Pr(A i)Pr(B|A i ) , (30.28)where the event A need not coincide with any of the A i .As a final point, we comment that sometimes we are concerned only with therelative probabilities of two events A <strong>and</strong> C (say), given the occurrence of someother event B. From (30.26) we then obtain a different <strong>for</strong>m of Bayes’ theorem,Pr(A|B)Pr(C|B) = Pr(A)Pr(B|A)Pr(C)Pr(B|C) , (30.29)which does not contain Pr(B) at all.30.3 Permutations <strong>and</strong> combinationsIn equation (30.5) we defined the probability of an event A in a sample space S asPr(A) = n A,n Swhere n A is the number of outcomes belonging to event A <strong>and</strong> n S is the totalnumber of possible outcomes. It is there<strong>for</strong>e necessary to be able to count thenumber of possible outcomes in various common situations.30.3.1 PermutationsLet us first consider a set of n objects that are all different. We may ask inhow many ways these n objects may be arranged, i.e. how many permutations ofthese objects exist. This is straight<strong>for</strong>ward to deduce, as follows: the object in thefirst position may be chosen in n different ways, that in the second position inn − 1 ways, <strong>and</strong> so on until the final object is positioned. The number of possiblearrangements is there<strong>for</strong>en(n − 1)(n − 2) ···(1) = n! (30.30)Generalising (30.30) slightly, let us suppose we choose only k (< n) objectsfrom n. The number of possible permutations of these k objects selected from nis given byn!n(n − 1)(n − 2) ···(n − k +1) =} {{ } (n − k)! ≡ n P k . (30.31)k factors1133


PROBABILITYIn calculating the number of permutations of the various objects we have sofar assumed that the objects are sampled without replacement – i.e. once an objecthas been drawn from the set it is put aside. As mentioned previously, however,we may instead replace each object be<strong>for</strong>e the next is chosen. The number ofpermutations of k objects from n with replacement may be calculated very easilysince the first object can be chosen in n different ways, as can the second, thethird, etc. There<strong>for</strong>e the number of permutations is simply n k . This may also beviewed as the number of permutations of k objects from n where repetitions areallowed, i.e. each object may be used as often as one likes.◮Find the probability that in a group of k people at least two have the same birthday(ignoring 29 February).It is simplest to begin by calculating the probability that no two people share a birthday,as follows. Firstly, we imagine each of the k people in turn pointing to their birthday ona year planner. Thus, we are sampling the 365 days of the year ‘with replacement’ <strong>and</strong> sothe total number of possible outcomes is (365) k . Now (<strong>for</strong> the moment) we assume that notwo people share a birthday <strong>and</strong> imagine the process being repeated, except that as eachperson points out their birthday it is crossed off the planner. In this case, we are samplingthe days of the year ‘without replacement’, <strong>and</strong> so the possible number of outcomes <strong>for</strong>which all the birthdays are different is365 365!P k =(365 − k)! .Hence the probability that all the birthdays are different is365!p =(365 − k)! 365 . kNow using the complement rule (30.11), the probability q that two or more people havethe same birthday is simply365!q =1− p =1−(365 − k)! 365 . kThis expression may be conveniently evaluated using Stirling’s approximation <strong>for</strong> n! whenn is large, namelyn! ∼ √ ( n) n2πn ,eto give( ) 365−k+0.5 365q ≈ 1 − e −k .365 − kIt is interesting to note that if k = 23 the probability is a little greater than a half thatat least two people have the same birthday, <strong>and</strong> if k = 50 the probability rises to 0.970.This can prove a good bet at a party of non-mathematicians! ◭So far we have assumed that all n objects are different (or distinguishable). Letus now consider n objects of which n 1 are identical <strong>and</strong> of type 1, n 2 are identical<strong>and</strong> of type 2, ...,n m are identical <strong>and</strong> of type m (clearly n = n 1 + n 2 + ···+ n m ).From (30.30) the number of permutations of these n objects is again n!. However,1134


30.3 PERMUTATIONS AND COMBINATIONSthe number of distinguishable permutations is onlyn!n 1 !n 2 ! ···n m ! , (30.32)since the ith group of identical objects can be rearranged in n i ! ways withoutchanging the distinguishable permutation.◮A set of snooker balls consists of a white, a yellow, a green, a brown, a blue, a pink, ablack <strong>and</strong> 15 reds. How many distinguishable permutations of the balls are there?In total there are 22 balls, the 15 reds being indistinguishable. Thus from (30.32) thenumber of distinguishable permutations is22!(1!)(1!)(1!)(1!)(1!)(1!)(1!)(15!) = 22! = 859 541 760. ◭15!30.3.2 CombinationsWe now consider the number of combinations of various objects when their orderis immaterial. Assuming all the objects to be distinguishable, from (30.31) we seethat the number of permutations of k objects chosen from n is n P k = n!/(n − k)!.Now, since we are no longer concerned with the order of the chosen objects, whichcan be internally arranged in k! different ways, the number of combinations of kobjects from n is( )n!n(n − k)!k! ≡ n C k ≡ <strong>for</strong> 0 ≤ k ≤ n, (30.33)kwhere, as noted in chapter 1, n C k is called the binomial coefficient since it alsoappears in the binomial expansion <strong>for</strong> positive integer n, namelyn∑(a + b) n =n C k a k b n−k . (30.34)k=0◮A h<strong>and</strong> of 13 playing cards is dealt from a well-shuffled pack of 52. What is the probabilitythat the h<strong>and</strong> contains two aces?Since the order of the cards in the h<strong>and</strong> is immaterial, the total number of distinct h<strong>and</strong>sis simply equal to the number of combinations of 13 objects drawn from 52, i.e. 52 C 13 .However, the number of h<strong>and</strong>s containing two aces is equal to the number of ways, 4 C 2 ,in which the two aces can be drawn from the four available, multiplied by the number ofways, 48 C 11 , in which the remaining 11 cards in the h<strong>and</strong> can be drawn from the 48 cardsthat are not aces. Thus the required probability is given by4 C 48 2 C 11= 4! 48! 13!39!52C 13 2!2! 11!37! 52!= (3)(4) (12)(13)(38)(39)2 (49)(50)(51)(52) =0.213 ◭1135


PROBABILITYAnother useful result that may be derived using the binomial coefficients is thenumber of ways in which n distinguishable objects can be divided into m piles,with n i objects in the ith pile, i =1, 2,...,m (the ordering of objects within eachpile being unimportant). This may be straight<strong>for</strong>wardly calculated as follows. Wemay choose the n 1 objects in the first pile from the original n objects in n C n1 ways.The n 2 objects in the second pile can then be chosen from the n − n 1 remainingobjects in n−n1 C n2 ways, etc. We may continue in this fashion until we reach the(m − 1)th pile, which may be <strong>for</strong>med in n−n1−···−nm−2 C nm−1 ways. The remainingobjects then <strong>for</strong>m the mth pile <strong>and</strong> so can only be ‘chosen’ in one way. Thus thetotal number of ways of dividing the original n objects into m piles is given bythe productN = n n−nC 1n1 C ···n−n1−···−nm−2 n2 C nm−1n! (n − n 1 )!=n 1 !(n − n 1 )! n 2 !(n − n 1 − n 2 )! ··· (n − n 1 − n 2 − ···− n m−2 )!n m−1 !(n − n 1 − n 2 − ···− n m−2 − n m−1 )!n! (n − n 1 )!=···(n − n 1 − n 2 − ···− n m−2 )!n 1 !(n − n 1 )! n 2 !(n − n 1 − n 2 )! n m−1 !n m !n!=n 1 !n 2 ! ···n m ! . (30.35)These numbers are called multinomial coefficients since (30.35) is the coefficient ofx n11 xn2 2 ···xnm m in the multinomial expansion of (x 1 + x 2 + ···+ x m ) n , i.e. <strong>for</strong> positiveinteger n∑(x 1 + x 2 + ···+ x m ) n n!=n 1 !n 2 ! ···n m ! xn1 1 xn2 2 ···xnm m .n 1 ,n 2 ,... ,nmn 1 +n 2 +···+nm=nFor the case m =2,n 1 = k, n 2 = n − k, (30.35) reduces to the binomial coefficientn C k . Furthermore, we note that the multinomial coefficient (30.35) is identical tothe expression (30.32) <strong>for</strong> the number of distinguishable permutations of n objects,n i of which are identical <strong>and</strong> of type i (<strong>for</strong> i =1, 2,...,m<strong>and</strong> n 1 +n 2 +···+n m = n).A few moments’ thought should convince the reader that the two expressions(30.35) <strong>and</strong> (30.32) must be identical.◮In the card game of bridge, each of four players is dealt 13 cards from a full pack of 52.What is the probability that each player is dealt an ace?From (30.35), the total number of distinct bridge dealings is 52!/(13!13!13!13!). However,the number of ways in which the four aces can be distributed with one in each h<strong>and</strong> is4!/(1!1!1!1!) = 4!; the remaining 48 cards can then be dealt out in 48!/(12!12!12!12!)ways. Thus the probability that each player receives an ace is4! 48! (13!) 4 24(13) 4=(12!) 4 52! (49)(50)(51)(52) =0.105. ◭As in the case of permutations we might ask how many combinations of kobjects can be chosen from n with replacement (repetition). To calculate this, we1136


30.3 PERMUTATIONS AND COMBINATIONSmay imagine the n (distinguishable) objects set out on a table. Each combinationof k objects can then be made by pointing to k of the n objectsinturn(withrepetitions allowed). These k equivalent selections distributed amongst n differentbut re-choosable objects are strictly analogous to the placing of k indistinguishable‘balls’ in n different boxes with no restriction on the number of balls in each box.A particular selection in the case k =7,n = 5 may be symbolised asxxx||x|xx|x.This denotes three balls in the first box, none in the second, one in the third, twoin the fourth <strong>and</strong> one in the fifth. We there<strong>for</strong>e need only consider the number of(distinguishable) ways in which k crosses <strong>and</strong> n − 1 vertical lines can be arranged,i.e. the number of permutations of k + n − 1 objects of which k are identicalcrosses <strong>and</strong> n − 1 are identical lines. This is given by (30.33) as(k + n − 1)!= n+k−1 C k . (30.36)k!(n − 1)!We note that this expression also occurs in the binomial expansion <strong>for</strong> negativeinteger powers. If n is a positive integer, it is straight<strong>for</strong>ward to show that (seechapter 1)∞∑(a + b) −n = (−1) k n+k−1 C k a −n−k b k ,k=0where a istakentobelargerthanb in magnitude.◮A system contains a number N of (non-interacting) particles, each of which can be inany of the quantum states of the system. The structure of the set of quantum states is suchthat there exist R energy levels with corresponding energies E i <strong>and</strong> degeneracies g i (i.e. theith energy level contains g i quantum states). Find the numbers of distinct ways in whichthe particles can be distributed among the quantum states of the system such that the ithenergy level contains n i particles, <strong>for</strong> i =1, 2,...,R, in the cases where the particles are(i) distinguishable with no restriction on the number in each state;(ii) indistinguishable with no restriction on the number in each state;(iii) indistinguishable with a maximum of one particle in each state;(iv) distinguishable with a maximum of one particle in each state.It is easiest to solve this problem in two stages. Let us first consider distributing the Nparticles among the R energy levels, without regard <strong>for</strong> the individual degenerate quantumstates that comprise each level. If the particles are distinguishable then the number ofdistinct arrangements with n i particles in the ith level, i =1, 2,...,R, is given by (30.35) asN!n 1 !n 2 ! ···n R ! .If, however, the particles are indistinguishable then clearly there exists only one distinctarrangement having n i particles in the ith level, i =1, 2,...,R . If we suppose that thereexist w i ways in which the n i particles in the ith energy level can be distributed amongthe g i degenerate states, then it follows that the number of distinct ways in which the N1137


PROBABILITYparticles can be distributed among all R quantum states of the system, with n i particles inthe ith level, is given by⎧N!R∏⎪⎨w i <strong>for</strong> distinguishable particles,n 1 !n 2 ! ···n R !i=1W {n i } = R∏⎪⎩ w i<strong>for</strong> indistinguishable particles.(30.37)i=1It there<strong>for</strong>e remains only <strong>for</strong> us to find the appropriate expression <strong>for</strong> w i in each of thecases (i)–(iv) above.Case (i). If there is no restriction on the number of particles in each quantum state,then in the ith energy level each particle can reside in any of the g i degenerate quantumstates. Thus, if the particles are distinguishable then the number of distinct arrangementsis simply w i = g n ii. Thus, from (30.37),W {n i } =N!n 1 !n 2 ! ···n R !R∏i=1g n ii = N!R∏i=1g n iin i ! .Such a system of particles (<strong>for</strong> example atoms or molecules in a classical gas) is said toobey Maxwell–Boltzmann statistics.Case (ii). If the particles are indistinguishable <strong>and</strong> there is no restriction on the numberin each state then, from (30.36), the number of distinct arrangements of the n i particlesamong the g i states in the ith energy level isw i = (n i + g i − 1)!n i !(g i − 1)! .Substituting this expression in (30.37), we obtainR∏ (n i + g i − 1)!W {n i } =n i !(g i − 1)! .i=1Such a system of particles (<strong>for</strong> example a gas of photons) is said to obey Bose–Einsteinstatistics.Case (iii). If a maximum of one particle can reside in each of the g i degenerate quantumstates in the ith energy level then the number of particles in each state is either 0 or 1.Since the particles are indistinguishable, w i is equal to the number of distinct arrangementsin which n i states are occupied <strong>and</strong> g i − n i states are unoccupied; this is given byThus, from (30.37), we havew i = g iC ni =g i !n i !(g i − n i )! .R∏ g i !W {n i } =n i !(g i − n i )! .i=1Such a system is said to obey Fermi–Dirac statistics, <strong>and</strong> an example is provided by anelectron gas.Case (iv). Again, the number of particles in each state is either 0 or 1. If the particlesare distinguishable, however, each arrangement identified in case (iii) can be reordered inn i ! different ways, so thatw i = g g i !iP ni =(g i − n i )! .1138


30.4 RANDOM VARIABLES AND DISTRIBUTIONSSubstituting this expression into (30.37) givesR∏ g i !W {n i } = N!n i !(g i − n i )! .i=1Such a system of particles has the names of no famous scientists attached to it, since itappears that it never occurs in nature. ◭30.4 R<strong>and</strong>om variables <strong>and</strong> distributionsSuppose an experiment has an outcome sample space S. A real variable X thatis defined <strong>for</strong> all possible outcomes in S (so that a real number – not necessarilyunique – is assigned to each possible outcome) is called a r<strong>and</strong>om variable (RV).The outcome of the experiment may already be a real number <strong>and</strong> hence a r<strong>and</strong>omvariable, e.g. the number of heads obtained in 10 throws of a coin, or the sum ofthe values if two dice are thrown. However, more arbitrary assignments are possible,e.g. the assignment of a ‘quality’ rating to each successive item produced by amanufacturing process. Furthermore, assuming that a probability can be assignedto all possible outcomes in a sample space S, it is possible to assign a probabilitydistribution to any r<strong>and</strong>om variable. R<strong>and</strong>om variables may be divided into twoclasses, discrete <strong>and</strong> continuous, <strong>and</strong> we now examine each of these in turn.30.4.1 Discrete r<strong>and</strong>om variablesA r<strong>and</strong>om variable X that takes only discrete values x 1 , x 2 , ...,x n , with probabilitiesp 1 , p 2 , ...,p n , is called a discrete r<strong>and</strong>om variable. The number of valuesn <strong>for</strong> which X has a non-zero probability is finite or at most countably infinite.As mentioned above, an example of a discrete r<strong>and</strong>om variable is the number ofheads obtained in 10 throws of a coin. If X is a discrete r<strong>and</strong>om variable, we c<strong>and</strong>efine a probability function (PF) f(x) that assigns probabilities to all the distinctvalues that X can take, such that{pi if x = x i ,f(x) =Pr(X = x) =(30.38)0 otherwise.A typical PF (see figure 30.6) thus consists of spikes, at valid values of X, whoseheight at x corresponds to the probability that X = x. Since the probabilitiesmust sum to unity, we requiren∑f(x i )=1. (30.39)i=1We may also define the cumulative probability function (CPF) of X, F(x), whosevalue gives the probability that X ≤ x, sothatF(x) =Pr(X ≤ x) = ∑ f(x i ). (30.40)x i≤x1139


PROBABILITYf(x)2pF(x)1p12 p 12 3 4 5 6x1 2 3 4 5 6(a)(b)Figure 30.6 (a) A typical probability function <strong>for</strong> a discrete distribution, that<strong>for</strong> the biased die discussed earlier. Since the probabilities must sum to unitywe require p =2/13. (b) The cumulative probability function <strong>for</strong> the samediscrete distribution. (Note that a different scale has been used <strong>for</strong> (b).)Hence F(x) is a step function that has upward jumps of p i at x = x i , i =1, 2,...,n, <strong>and</strong> is constant between possible values of X. We may also calculatethe probability that X lies between two limits, l 1 <strong>and</strong> l 2 (l 1


30.4 RANDOM VARIABLES AND DISTRIBUTIONSf(x)l 1 a b l 2xFigure 30.7 The probability density function <strong>for</strong> a continuous r<strong>and</strong>om variableX that can take values only between the limits l 1 <strong>and</strong> l 2 . The shaded areaunder the curve gives Pr(a


PROBABILITY◮A r<strong>and</strong>om variable X has a PDF f(x) given by Ae −x in the interval 0


30.5 PROPERTIES OF DISTRIBUTIONSIn many circumstances, however, r<strong>and</strong>om variables do not depend on oneanother, i.e. they are independent. As an example, <strong>for</strong> a person drawn at r<strong>and</strong>omfrom a population, we might expect height <strong>and</strong> IQ to be independent r<strong>and</strong>omvariables. Let us suppose that X <strong>and</strong> Y are two r<strong>and</strong>om variables with probabilitydensity functions g(x) <strong>and</strong>h(y) respectively. In mathematical terms, X <strong>and</strong> Y areindependent RVs if their joint probability density function is given by f(x, y) =g(x)h(y). Thus, <strong>for</strong> independent RVs, if X <strong>and</strong> Y are both discrete thenPr(X = x i , Y = y j )=g(x i )h(y j )or, if X <strong>and</strong> Y are both continuous, thenPr(x


PROBABILITYthe series is absolutely convergent or that the integral exists, as the case may be.From its definition it is straight<strong>for</strong>ward to show that the expectation value hasthe following properties:(i) if a is a constant then E[a] =a;(ii) if a is a constant then E[ag(X)] = aE[g(X)];(iii) if g(X) =s(X)+t(X) thenE[ g(X)] = E[ s(X)] + E[t(X)].It should be noted that the expectation value is not a function of X but isinstead a number that depends on the <strong>for</strong>m of the probability density functionf(x) <strong>and</strong> the function g(x). Most of the st<strong>and</strong>ard quantities used to characterisef(x) are simply the expectation values of various functions of the r<strong>and</strong>om variableX. We now consider these st<strong>and</strong>ard quantities.30.5.1 MeanThe property most commonly used to characterise a probability distribution isits mean, which is defined simply as the expectation value E[X] of the variable Xitself. Thus, the mean is given by{∑iE[X] =x if(x i ) <strong>for</strong> a discrete distribution,∫ (30.46)xf(x) dx <strong>for</strong> a continuous distribution.The alternative notations µ <strong>and</strong> 〈x〉 are also commonly used to denote the mean.If in (30.46) the series is not absolutely convergent, or the integral does not exist,we say that the distribution does not have a mean, but this is very rare in physicalapplications.◮The probability of finding a 1s electron in a hydrogen atom in a given infinitesimal volumedV is ψ ∗ ψdV, where the quantum mechanical wavefunction ψ is given byψ = Ae −r/a 0.Find the value of the real constant A <strong>and</strong> thereby deduce the mean distance of the electronfrom the origin.Let us consider the r<strong>and</strong>om variable R = ‘distance of the electron from the origin’. Sincethe 1s orbital has no θ- orφ-dependence (it is spherically symmetric), we may considerthe infinitesimal volume element dV as the spherical shell with inner radius r <strong>and</strong> outerradius r + dr. Thus, dV =4πr 2 dr <strong>and</strong> the PDF of R is simplyPr(r


30.5 PROPERTIES OF DISTRIBUTIONSIntegrating by parts we find A =1/(πa 3 0 )1/2 . Now, using the definition of the mean (30.46),we find∫ ∞E[R] = rf(r) dr = 4 r 3 e −2r/a 0dr.0a 3 0 0The integral on the RHS may be integrated by parts <strong>and</strong> takes the value 3a 4 0 /8; consequentlywe find that E[R] =3a 0 /2. ◭∫ ∞30.5.2 Mode <strong>and</strong> medianAlthough the mean discussed in the last section is the most common measureof the ‘average’ of a distribution, two other measures, which do not rely on theconcept of expectation values, are frequently encountered.The mode of a distribution is the value of the r<strong>and</strong>om variable X at which theprobability (density) function f(x) has its greatest value. If there is more than onevalue of X <strong>for</strong> which this is true then each value may equally be called the modeof the distribution.The median M of a distribution is the value of the r<strong>and</strong>om variable X at whichthe cumulative probability function F(x) takes the value 1 2 ,i.e.F(M) = 1 2 . Relatedto the median are the lower <strong>and</strong> upper quartiles Q l <strong>and</strong> Q u of the PDF, whichare defined such thatF(Q l )= 1 4 , F(Q u)= 3 4 .Thus the median <strong>and</strong> lower <strong>and</strong> upper quartiles divide the PDF into four regionseach containing one quarter of the probability. Smaller subdivisions are alsopossible, e.g. the nth percentile, P n , of a PDF is defined by F(P n )=n/100.◮Find the mode of the PDF <strong>for</strong> the distance from the origin of the electron whose wavefunctionwas given in the previous example.We found in the previous example that the PDF <strong>for</strong> the electron’s distance from the originwas given byf(r) = 4r2 e −2r/a 0. (30.47)a 3 0Differentiating f(r) with respect to r, weobtaindfdr = 8ra 3 0(1 − r a 0)e −2r/a 0.Thus f(r) has turning points at r =0<strong>and</strong>r = a 0 ,wheredf/dr = 0. It is straight<strong>for</strong>wardto show that r = 0 is a minimum <strong>and</strong> r = a 0 is a maximum. Moreover, it is also clear thatr = a 0 is a global maximum (as opposed to just a local one). Thus the mode of f(r) occursat r = a 0 . ◭1145


PROBABILITY30.5.3 Variance <strong>and</strong> st<strong>and</strong>ard deviationThe variance of a distribution, V [X], also written σ 2 , is defined byV [X] =E [ {∑(X − µ) 2] j=(x j − µ) 2 f(x j )∫(x − µ) 2 f(x) dx<strong>for</strong> a discrete distribution,<strong>for</strong> a continuous distribution.(30.48)Here µ has been written <strong>for</strong> the expectation value E[X] ofX. As in the case ofthe mean, unless the series <strong>and</strong> the integral in (30.48) converge the distributiondoes not have a variance. From the definition (30.48) we may easily derive thefollowing useful properties of V [X]. If a <strong>and</strong> b are constants then(i) V [a] =0,(ii) V [aX + b] =a 2 V [X].The variance of a distribution is always positive; its positive square root isknown as the st<strong>and</strong>ard deviation of the distribution <strong>and</strong> is often denoted by σ.Roughly speaking, σ measures the spread (about x = µ) of the values that X canassume.◮Find the st<strong>and</strong>ard deviation of the PDF <strong>for</strong> the distance from the origin of the electronwhose wavefunction was discussed in the previous two examples.Inserting the expression (30.47) <strong>for</strong> the PDF f(r) into (30.48), the variance of the r<strong>and</strong>omvariable R is given by∫ ∞V [R] = (r − µ) 2 4r2 e −2r/a 0dr = 4 (r 4 − 2r 3 µ + r 2 µ 2 )e −2r/a 0dr,0 a 3 0a 3 0 0where the mean µ = E[R] =3a 0 /2. Integrating each term in the integr<strong>and</strong> by parts weobtain∫ ∞V [R] =3a 2 0 − 3µa 0 + µ 2 = 3a2 04 .Thus the st<strong>and</strong>ard deviation of the distribution is σ = √ 3a 0 /2. ◭We may also use the definition (30.48) to derive the Bienaymé–Chebyshevinequality, which provides a useful upper limit on the probability that r<strong>and</strong>omvariable X takes values outside a given range centred on the mean. Let us considerthe case of a continuous r<strong>and</strong>om variable, <strong>for</strong> which∫Pr(|X − µ| ≥c) = f(x) dx,|x−µ|≥cwhere the integral on the RHS extends over all values of x satisfying the inequality1146


30.5 PROPERTIES OF DISTRIBUTIONS|x − µ| ≥c. From (30.48), we find that∫∫σ 2 ≥ (x − µ) 2 f(x) dx ≥ c 2|x−µ|≥c|x−µ|≥cf(x) dx. (30.49)The first inequality holds because both (x − µ) 2 <strong>and</strong> f(x) are non-negative <strong>for</strong>all x, <strong>and</strong> the second inequality holds because (x − µ) 2 ≥ c 2 over the range ofintegration. However, the RHS of (30.49) is simply equal to c 2 Pr(|X − µ| ≥c),<strong>and</strong> thus we obtain the required inequalityPr(|X − µ| ≥c) ≤ σ2c 2 .A similar derivation may be carried through <strong>for</strong> the case of a discrete r<strong>and</strong>omvariable. Thus, <strong>for</strong> any distribution f(x) that possesses a variance we have, <strong>for</strong>example,Pr(|X − µ| ≥2σ) ≤ 1 4<strong>and</strong> Pr(|X − µ| ≥3σ) ≤ 1 9 .30.5.4 MomentsThe mean (or expectation) of X is sometimes called the first moment of X, sinceit is defined as the sum or integral of the probability density function multipliedby the first power of x. By a simple extension the kth moment of a distributionis defined by{∑µ k ≡ E[X k j]=xk j f(x j) <strong>for</strong> a discrete distribution,∫x k f(x) dx <strong>for</strong> a continuous distribution.(30.50)For notational convenience, we have introduced the symbol µ k to denote E[X k ],the kth moment of the distribution. Clearly, the mean of the distribution is thendenoted by µ 1 , often abbreviated simply to µ, as in the previous subsection, asthis rarely causes confusion.A useful result that relates the second moment, the mean <strong>and</strong> the variance ofa distribution is proved using the properties of the expectation operator:V [X] =E [ (X − µ) 2]= E [ X 2 − 2µX + µ 2]= E [ X 2] − 2µE[X]+µ 2= E [ X 2] − 2µ 2 + µ 2In alternative notations, this result can be written= E [ X 2] − µ 2 . (30.51)〈(x − µ) 2 〉 = 〈x 2 〉−〈x〉 2 or σ 2 = µ 2 − µ 2 1.1147


PROBABILITY◮A biased die has probabilities p/2, p, p, p, p, 2p of showing 1, 2, 3, 4, 5, 6 respectively. Find(i) the mean, (ii) the second moment <strong>and</strong> (iii) the variance of this probability distribution.By dem<strong>and</strong>ing that the sum of the probabilities equals unity we require p =2/13. Now,using the definition of the mean (30.46) <strong>for</strong> a discrete distribution,E[X] = ∑ jx j f(x j )=1× 1 p + 2× p + 3× p + 4× p + 5× p + 6× 2p2= 53 2 p = 532 × 213 = 5313 .Similarly, using the definition of the second moment (30.50),E[X 2 ]= ∑ x 2 j f(x j )=1 2 × 1 p 2 +22 p +3 2 p +4 2 p +5 2 p +6 2 × 2pj= 2532 p = 25313 .Finally, using the definition of the variance (30.48), with µ =53/13, we obtainV [X] = ∑ (x j − µ) 2 f(x j )j=(1− µ) 2 1 p +(2− 2 µ)2 p +(3− µ) 2 p +(4− µ) 2 p +(5− µ) 2 p +(6− µ) 2 2p( ) 3120= p = 480169 169 .It is easy to verify that V [X] =E [ X 2] − (E[X]) 2 . ◭In practice, to calculate the moments of a distribution it is often simpler to usethe moment generating function discussed in subsection 30.7.2. This is particularlytrue <strong>for</strong> higher-order moments, where direct evaluation of the sum or integral in(30.50) can be somewhat laborious.30.5.5 Central momentsThe variance V [X] is sometimes called the second central moment of the distribution,since it is defined as the sum or integral of the probability density functionmultiplied by the second power of x − µ. The origin of the term ‘central’ is that bysubtracting µ from x be<strong>for</strong>e squaring we are considering the moment about themean of the distribution, rather than about x = 0. Thus the kth central momentof a distribution is defined asν k ≡ E [ {∑(X − µ) k] j=(x j − µ) k f(x j ) <strong>for</strong> a discrete distribution,∫ (30.52)(x − µ) k f(x) dx <strong>for</strong> a continuous distribution.It is convenient to introduce the notation ν k <strong>for</strong> the kth central moment. ThusV [X] ≡ ν 2 <strong>and</strong> we may write (30.51) as ν 2 = µ 2 − µ 2 1 . Clearly, the first centralmoment of a distribution is always zero since, <strong>for</strong> example in the continuous case,∫∫∫ν 1 = (x − µ)f(x) dx = xf(x) dx − µ f(x) dx = µ − (µ × 1) = 0.1148


30.5 PROPERTIES OF DISTRIBUTIONSWe note that the notation µ k <strong>and</strong> ν k <strong>for</strong> the moments <strong>and</strong> central momentsrespectively is not universal. Indeed, in some books their meanings are reversed.We can write the kth central moment of a distribution in terms of its kth <strong>and</strong>lower-order moments by exp<strong>and</strong>ing (X − µ) k in powers of X. We have alreadynoted that ν 2 = µ 2 − µ 2 1 , <strong>and</strong> similar expressions may be obtained <strong>for</strong> higher-ordercentral moments. For example,In general, it is straight<strong>for</strong>ward to show thatν 3 = E [ (X − µ 1 ) 3]= E [ X 3 − 3µ 1 X 2 +3µ 2 1X − µ 3 ]1= µ 3 − 3µ 1 µ 2 +3µ 2 1µ 1 − µ 3 1= µ 3 − 3µ 1 µ 2 +2µ 3 1. (30.53)ν k = µ k − k C 1 µ k−1 µ 1 + ···+(−1) r k C r µ k−r µ r 1 + ···+(−1) k−1 ( k C k−1 − 1)µ k 1.(30.54)Once again, direct evaluation of the sum or integral in (30.52) can be rathertedious <strong>for</strong> higher moments, <strong>and</strong> it is usually quicker to use the moment generatingfunction (see subsection 30.7.2), from which the central moments can be easilyevaluated as well.◮The PDF <strong>for</strong> a Gaussian distribution (see subsection 30.9.1) with mean µ <strong>and</strong> varianceσ 2 is given byf(x) = 1 [ ]σ √ 2π exp (x − µ)2− .2σ 2Obtain an expression <strong>for</strong> the kth central moment of this distribution.As an illustration, we will per<strong>for</strong>m this calculation by evaluating the integral in (30.52)directly. Thus, the kth central moment of f(x) isgivenbyν k =∫ ∞−∞= 1σ √ 2π= 1σ √ 2π(x − µ) k f(x) dx∫ ∞−∞∫ ∞−∞](x − µ) k (x − µ)2exp[−2σ 2dx)y k exp(− y2dy, (30.55)2σ 2where in the last line we have made the substitution y = x − µ. It is clear that if k isodd then the integr<strong>and</strong> is an odd function of y <strong>and</strong> hence the integral equals zero. Thus,ν k =0ifk is odd. When k is even, we could calculate ν k by integrating by parts to obtaina reduction <strong>for</strong>mula, but it is more elegant to consider instead the st<strong>and</strong>ard integral (seesubsection 6.4.2)I =∫ ∞−∞exp(−αy 2 ) dy = π 1/2 α −1/2 ,1149


PROBABILITY<strong>and</strong> differentiate it repeatedly with respect to α (see section 5.12). Thus, we obtain∫dI∞dα = − y 2 exp(−αy 2 ) dy = − 1 2 π1/2 α −3/2−∞d 2 ∫I ∞dα = y 4 exp(−αy 2 ) dy =( 1 )( 3 2 2 2 )π1/2 α −5/2−∞.d n ∫I∞dα n =(−1)n y 2n exp(−αy 2 ) dy =(−1) n ( 1 )( 3 ) ···( 1 (2n − 2 2 2 1))π1/2 α −(2n+1)/2 .−∞Setting α =1/(2σ 2 ) <strong>and</strong> substituting the above result into (30.55), we find (<strong>for</strong> k even)ν k =( 1 )( 3 ) ···( 1 (k − 2 2 2 1))(2σ2 ) k/2 = (1)(3) ···(k − 1)σ k . ◭One may also characterise a probability distribution f(x) using the closelyrelated normalised <strong>and</strong> dimensionless central momentsγ k ≡ν k= ν kν k/2 σ k .2From this set, γ 3 <strong>and</strong> γ 4 are more commonly called, respectively, the skewness<strong>and</strong> kurtosis of the distribution. The skewness γ 3 of a distribution is zero if it issymmetrical about its mean. If the distribution is skewed to values of x smallerthan the mean then γ 3 < 0. Similarly γ 3 > 0 if the distribution is skewed to highervalues of x.From the above example, we see that the kurtosis of the Gaussian distribution(subsection 30.9.1) is given byγ 4 = ν 4ν22 = 3σ4σ 4 =3.It is there<strong>for</strong>e common practice to define the excess kurtosis of a distributionas γ 4 − 3. A positive value of the excess kurtosis implies a relatively narrowerpeak <strong>and</strong> wider wings than the Gaussian distribution with the same mean <strong>and</strong>variance. A negative excess kurtosis implies a wider peak <strong>and</strong> shorter wings.Finally, we note here that one can also describe a probability density functionf(x) in terms of its cumulants, which are again related to the central moments.However, we defer the discussion of cumulants until subsection 30.7.4, since theirdefinition is most easily understood in terms of generating functions.30.6 Functions of r<strong>and</strong>om variablesSuppose X is some r<strong>and</strong>om variable <strong>for</strong> which the probability density functionf(x) is known. In many cases, we are more interested in a related r<strong>and</strong>om variableY = Y (X), where Y (X) is some function of X. What is the probability density1150


30.6 FUNCTIONS OF RANDOM VARIABLESfunction g(y) <strong>for</strong> the new r<strong>and</strong>om variable Y ? We now discuss how to obtainthis function.30.6.1 Discrete r<strong>and</strong>om variablesIf X is a discrete RV that takes only the values x i , i =1, 2,...,n,thenY mustalso be discrete <strong>and</strong> takes the values y i = Y (x i ), although some of these valuesmay be identical. The probability function <strong>for</strong> Y is given by{∑jg(y) =f(x j) if y = y i ,(30.56)0 otherwise,where the sum extends over those values of j <strong>for</strong> which y i = Y (x j ). The simplestcase arises when the function Y (X) possesses a single-valued inverse X(Y ). In thiscase, only one x-value corresponds to each y-value, <strong>and</strong> we obtain a closed-<strong>for</strong>mexpression <strong>for</strong> g(y) given by{f(x(yi )) if y = y i ,g(y) =0 otherwise.If Y (X) does not possess a single-valued inverse then the situation is morecomplicated <strong>and</strong> it may not be possible to obtain a closed-<strong>for</strong>m expression <strong>for</strong>g(y). Nevertheless, whatever the <strong>for</strong>m of Y (X), one can always use (30.56) toobtain the numerical values of the probability function g(y) aty = y i .30.6.2 Continuous r<strong>and</strong>om variablesIf X is a continuous RV, then so too is the new r<strong>and</strong>om variable Y = Y (X). Theprobability that Y lies in the range y to y + dy is given by∫g(y) dy = f(x) dx, (30.57)where dS corresponds to all values of x <strong>for</strong> which Y lies in the range y to y + dy.Once again the simplest case occurs when Y (X) possesses a single-valued inverseX(Y ). In this case, we may writeg(y) dy =∣∫ x(y+dy)x(y)dS∣ ∫ ∣∣∣ x(y)+| dxf(x ′ ) dx ′ =x(y)dy | dyf(x ′ ) dx ′ ,from which we obtaing(y) =f(x(y))dx∣ dy ∣ . (30.58)1151


PROBABILITYθlighthouseLbeamOcoastlineyFigure 30.8The illumination of a coastline by the beam from a lighthouse.◮A lighthouse is situated at a distance L from a straight coastline, opposite a point O, <strong>and</strong>sends out a narrow continuous beam of light simultaneously in opposite directions. The beamrotates with constant angular velocity. If the r<strong>and</strong>om variable Y is the distance along thecoastline, measured from O, of the spot that the light beam illuminates, find its probabilitydensity function.The situation is illustrated in figure 30.8. Since the light beam rotates at a constant angularvelocity, θ is distributed uni<strong>for</strong>mly between −π/2 <strong>and</strong> π/2, <strong>and</strong> so f(θ) =1/π. Nowy = L tan θ, which possesses the single-valued inverse θ =tan −1 (y/L), provided that θ liesbetween −π/2 <strong>and</strong>π/2. Since dy/dθ = L sec 2 θ = L(1 + tan 2 θ)=L[1 + (y/L) 2 ], from(30.58) we findg(y) = 1 dθπ ∣ dy ∣ = 1<strong>for</strong> −∞


30.6 FUNCTIONS OF RANDOM VARIABLESYy + dyydx 1 dx 2XFigure 30.9 Illustration of a function Y (X) whose inverse X(Y ) is a doublevaluedfunction of Y . The range y to y + dy corresponds to X being either inthe range x 1 to x 1 + dx 1 or in the range x 2 to x 2 + dx 2 .This result may be generalised straight<strong>for</strong>wardly to the case where the range y toy + dy corresponds to more than two x-intervals.◮The r<strong>and</strong>om variable X is Gaussian distributed (see subsection 30.9.1) with mean µ <strong>and</strong>variance σ 2 . Find the PDF of the new variable Y =(X − µ) 2 /σ 2 .It is clear that X(Y ) is a double-valued function of Y . However, in this case, it isstraight<strong>for</strong>ward to obtain single-valued functions giving the two values of x that correspondto a given value of y; thesearex 1 = µ − σ √ y <strong>and</strong> x 2 = µ + σ √ y,where √ y is taken tomean the positive square root.The PDF of X is given byf(x) = 1 [ ]σ √ 2π exp (x − µ)2− .2σ 2Since dx 1 /dy = −σ/(2 √ y)<strong>and</strong>dx 2 /dy = σ/(2 √ y), from (30.59) we obtaing(y) = 1∣ ∣∣∣σ √ 2π exp(− 1 y) −σ22 √ y ∣ + 1∣ ∣∣∣σ √ 2π exp(− 1 y) σ22 √ y ∣= 12 √ π ( 1 2 y)−1/2 exp(− 1 y). 2As we shall see in subsection 30.9.3, this is the gamma distribution γ( 1 , 1 ). ◭ 2 230.6.3 Functions of several r<strong>and</strong>om variablesWe may extend our discussion further, to the case in which the new r<strong>and</strong>omvariable is a function of several other r<strong>and</strong>om variables. For definiteness, let usconsider the r<strong>and</strong>om variable Z = Z(X,Y ), which is a function of two otherRVs X <strong>and</strong> Y . Given that these variables are described by the joint probabilitydensity function f(x, y), we wish to find the probability density function p(z) ofthe variable Z.1153


PROBABILITYIf X <strong>and</strong> Y are both discrete RVs thenp(z) = ∑ i,jf(x i ,y j ), (30.60)where the sum extends over all values of i <strong>and</strong> j <strong>for</strong> which Z(x i ,y j )=z. Similarly,if X <strong>and</strong> Y are both continuous RVs then p(z) is found by requiring that∫∫p(z) dz = f(x, y) dx dy, (30.61)dSwhere dS is the infinitesimal area in the xy-plane lying between the curvesZ(x, y) =z <strong>and</strong> Z(x, y) =z + dz.◮Suppose X <strong>and</strong> Y are independent continuous r<strong>and</strong>om variables in the range −∞ to ∞,with PDFs g(x) <strong>and</strong> h(y) respectively. Obtain expressions <strong>for</strong> the PDFs of Z = X + Y <strong>and</strong>W = XY .Since X <strong>and</strong> Y are independent RVs, their joint PDF is simply f(x, y) =g(x)h(y). Thus,from (30.61), the PDF of the sum Z = X + Y is given by∫ ∞∫ z+dz−xp(z) dz = dx g(x) dy h(y)−∞z−x(∫ ∞)= g(x)h(z − x) dx dz.−∞Thus p(z) istheconvolution of the PDFs of g <strong>and</strong> h (i.e. p = g ∗ h, see subsection 13.1.7).In a similar way, the PDF of the product W = XY is given by∫ ∞∫ (w+dw)/|x|q(w) dw = dx g(x) dy h(y)−∞w/|x|(∫ ∞= g(x)h(w/x) dx )dw ◭−∞|x|The prescription (30.61) is readily generalised to functions of n r<strong>and</strong>om variablesZ = Z(X 1 ,X 2 ,...,X n ), in which case the infinitesimal ‘volume’ element dS is theregion in x 1 x 2 ···x n -space between the (hyper)surfaces Z(x 1 ,x 2 ,...,x n )=z <strong>and</strong>Z(x 1 ,x 2 ,...,x n )=z + dz. In practice, however, the integral is difficult to evaluate,since one is faced with the complicated geometrical problem of determining thelimits of integration. Fortunately, an alternative (<strong>and</strong> powerful) technique exists<strong>for</strong> evaluating integrals of this kind. One eliminates the geometrical problem byintegrating over all values of the variables x i without restriction, while shiftingthe constraint on the variables to the integr<strong>and</strong>. This is readily achieved bymultiplying the integr<strong>and</strong> by a function that equals unity in the infinitesimalregion dS <strong>and</strong> zero elsewhere. From the discussion of the Dirac delta function insubsection 13.1.3, we see that δ(Z(x 1 ,x 2 ,...,x n )−z) dz satisfies these requirements,<strong>and</strong> so in the most general case we have∫∫ ∫p(z) = ··· f(x 1 ,x 2 ,...,x n )δ(Z(x 1 ,x 2 ,...,x n ) − z) dx 1 dx 2 ...dx n ,(30.62)1154


30.6 FUNCTIONS OF RANDOM VARIABLESwhere the range of integration is over all possible values of the variables x i .Thisintegral is most readily evaluated by substituting in (30.62) the Fourier integralrepresentation of the Dirac delta function discussed in subsection 13.1.4, namelyδ(Z(x 1 ,x 2 ,...,x n ) − z) = 1 ∫ ∞e ik(Z(x1,x2,...,xn)−z) dk. (30.63)2π −∞This is best illustrated by considering a specific example.◮A general one-dimensional r<strong>and</strong>om walk consists of n independent steps, each of whichcan be of a different length <strong>and</strong> in either direction along the x-axis. If g(x) is the PDF <strong>for</strong>the (positive or negative) displacement X along the x-axis achieved in a single step, obtainan expression <strong>for</strong> the PDF of the total displacement S after n steps.The total displacement S is simply the algebraic sum of the displacements X i achieved ineach of the n steps, so thatS = X 1 + X 2 + ···+ X n .Since the r<strong>and</strong>om variables X i are independent <strong>and</strong> have the same PDF g(x), their jointPDF is simply g(x 1 )g(x 2 ) ···g(x n ). Substituting this into (30.62), together with (30.63), weobtain∫ ∞ ∫ ∞ ∫ ∞p(s) = ··· g(x 1 )g(x 2 ) ···g(x n ) 1 ∫ ∞e ik[(x 1+x 2 +···+x n)−s] dk dx 1 dx 2 ···dx n2π= 12π−∞ −∞∫ ∞−∞−∞−∞(∫ ∞ndk e −iks g(x)e dx) ikx . (30.64)−∞It is convenient to define the characteristic function C(k) of the variable X asC(k) =∫ ∞−∞g(x)e ikx dx,which is simply related to the Fourier trans<strong>for</strong>m of g(x). Then (30.64) may be written asp(s) = 1 ∫ ∞e −iks [C(k)] n dk.2π −∞Thus p(s) can be found by evaluating two Fourier integrals. Characteristic functions willbe discussed in more detail in subsection 30.7.3. ◭30.6.4 Expectation values <strong>and</strong> variancesIn some cases, one is interested only in the expectation value or the varianceof the new variable Z rather than in its full probability density function. Fordefiniteness, let us consider the r<strong>and</strong>om variable Z = Z(X,Y ), which is a functionof two RVs X <strong>and</strong> Y with a known joint distribution f(x, y); the results we willobtain are readily generalised to more (or fewer) variables.It is clear that E[Z] <strong>and</strong>V [Z] can be obtained, in principle, by first using themethods discussed above to obtain p(z) <strong>and</strong> then evaluating the appropriate sumsor integrals. The intermediate step of calculating p(z) is not necessary, however,since it is straight<strong>for</strong>ward to obtain expressions <strong>for</strong> E[Z] <strong>and</strong>V [Z] in terms of1155


PROBABILITYthe variables X <strong>and</strong> Y . For example, if X <strong>and</strong> Y are continuous RVs then theexpectation value of Z is given by∫∫∫E[Z] = zp(z) dz = Z(x, y)f(x, y) dx dy. (30.65)An analogous result exists <strong>for</strong> discrete r<strong>and</strong>om variables.Integrals of the <strong>for</strong>m (30.65) are often difficult to evaluate. Nevertheless, wemay use (30.65) to derive an important general result concerning expectationvalues. If X <strong>and</strong> Y are any two r<strong>and</strong>om variables <strong>and</strong> a <strong>and</strong> b are arbitraryconstants then by letting Z = aX + bY we findE[aX + bY ]=aE[X]+bE[Y ].Furthermore, we may use this result to obtain an approximate expression <strong>for</strong> theexpectation value E[ Z(X,Y )] of any arbitrary function of X <strong>and</strong> Y . Letting µ X =E[X] <strong>and</strong>µ Y = E[Y ], <strong>and</strong> provided Z(X,Y ) can be reasonably approximated bythe linear terms of its Taylor expansion about the point (µ X ,µ Y ), we haveZ(X,Y ) ≈ Z(µ X ,µ Y )+( ∂Z∂X)(X − µ X )+( ∂Z∂Y)(Y − µ Y ),(30.66)where the partial derivatives are evaluated at X = µ X <strong>and</strong> Y = µ Y . Taking theexpectation values of both sides, we find( )( )∂Z∂ZE[ Z(X,Y )] ≈ Z(µ X ,µ Y )+ (E[X]−µ X )+ (E[Y ]−µ Y )=Z(µ X ,µ Y ),∂X∂Ywhich gives the approximate result E[ Z(X,Y )] ≈ Z(µ X ,µ Y ).By analogy with (30.65), the variance of Z = Z(X,Y ) is given by∫∫∫V [Z] = (z − µ Z ) 2 p(z) dz = [Z(x, y) − µ Z ] 2 f(x, y) dx dy,(30.67)where µ Z = E[Z]. We may use this expression to derive a second useful result. IfX <strong>and</strong> Y are two independent r<strong>and</strong>om variables, so that f(x, y) =g(x)h(y), <strong>and</strong>a, b <strong>and</strong> c are constants then by setting Z = aX + bY + c in (30.67) we obtainV [aX + bY + c] =a 2 V [X]+b 2 V [Y ]. (30.68)From (30.68) we also obtain the important special caseV [X + Y ]=V [X − Y ]=V [X]+V [Y ].Provided X <strong>and</strong> Y are indeed independent r<strong>and</strong>om variables, we may obtainan approximate expression <strong>for</strong> V [ Z(X,Y )], <strong>for</strong> any arbitrary function Z(X,Y ),in a similar manner to that used in approximating E[ Z(X,Y )] above. Taking the1156


30.7 GENERATING FUNCTIONSvariance of both sides of (30.66), <strong>and</strong> using (30.68), we find( ) 2 ( ) 2 ∂Z∂ZV [ Z(X,Y )] ≈ V [X]+ V [Y ], (30.69)∂X∂Ythe partial derivatives being evaluated at X = µ X <strong>and</strong> Y = µ Y .30.7 Generating functionsAs we saw in chapter 16, when dealing with particular sets of functions f n ,each member of the set being characterised by a different non-negative integern, it is sometimes possible to summarise the whole set by a single function of adummy variable (say t), called a generating function. The relationship betweenthe generating function <strong>and</strong> the nth member f n of the set is that if the generatingfunction is exp<strong>and</strong>ed as a power series in t then f n is the coefficient of t n .Forexample, in the expansion of the generating function G(z,t) =(1− 2zt + t 2 ) −1/2 ,the coefficient of t n is the nth Legendre polynomial P n (z), i.e.∞∑G(z,t) =(1− 2zt + t 2 ) −1/2 = P n (z)t n .We found that many useful properties of, <strong>and</strong> relationships between, the membersof a set of functions could be established using the generating function <strong>and</strong> otherfunctions obtained from it, e.g. its derivatives.Similar ideas can be used in the area of probability theory, <strong>and</strong> two types ofgenerating function can be usefully defined, one more generally applicable thanthe other. The more restricted of the two, applicable only to discrete integraldistributions, is called a probability generating function; this is discussed in thenext section. The second type, a moment generating function, can be used withboth discrete <strong>and</strong> continuous distributions <strong>and</strong> is considered in subsection 30.7.2.From the moment generating function, we may also construct the closely relatedcharacteristic <strong>and</strong> cumulant generating functions; these are discussed insubsections 30.7.3 <strong>and</strong> 30.7.4 respectively.n=030.7.1 Probability generating functionsAs already indicated, probability generating functions are restricted in applicabilityto integer distributions, of which the most common (the binomial, the Poisson<strong>and</strong> the geometric) are considered in this <strong>and</strong> later subsections. In such distributionsa r<strong>and</strong>om variable may take only non-negative integer values. The actualpossible values may be finite or infinite in number, but, <strong>for</strong> <strong>for</strong>mal purposes,all integers, 0, 1, 2,... are considered possible. If only a finite number of integervalues can occur in any particular case then those that cannot occur are includedbut are assigned zero probability.1157


PROBABILITYIf, as previously, the probability that the r<strong>and</strong>om variable X takes the value x nis f(x n ), then∑f(x n )=1.nIn the present case, however, only non-negative integer values of x n are possible,<strong>and</strong> we can, without ambiguity, write the probability that X takes the value n asf n , with∞∑f n =1. (30.70)We may now define the probability generating function Φ X (t) byn=0Φ X (t) ≡∞∑f n t n . (30.71)It is immediately apparent that Φ X (t) =E[t X ] <strong>and</strong> that, by virtue of (30.70),Φ X (1) = 1.Probably the simplest example of a probability generating function (PGF) isprovided by the r<strong>and</strong>om variable X defined by{1 if the outcome of a single trial is a ‘success’,X =0 if the trial ends in ‘failure’.If the probability of success is p <strong>and</strong> that of failure q (= 1 − p) thenn=0Φ X (t) =qt 0 + pt 1 +0+0+···= q + pt. (30.72)This type of r<strong>and</strong>om variable is discussed much more fully in subsection 30.8.1.In a similar but slightly more complicated way, a Poisson-distributed integervariable with mean λ (see subsection 30.8.4) has a PGFΦ X (t) =∞∑n=0e −λ λ nt n = e −λ e λt . (30.73)n!We note that, as required, Φ X (1) = 1 in both cases.Useful results will be obtained from this kind of approach only if the summation(30.71) can be carried out explicitly in particular cases <strong>and</strong> the functions derivedfrom Φ X (t) can be shown to be related to meaningful parameters. Two suchrelationships can be obtained by differentiating (30.71) with respect to t. Takingthe first derivative we finddΦ X (t)dt∞∑∞∑= nf n t n−1 ⇒ Φ ′ X(1) = nf n = E[X], (30.74)n=01158n=0


30.7 GENERATING FUNCTIONS<strong>and</strong> differentiating once more we obtaind 2 Φ X (t)∞∑dt 2 = n(n − 1)f n t n−2 ⇒ Φ ′′ X (1) = ∑∞ n(n − 1)f n = E[X(X − 1)].n=0n=0(30.75)Equation (30.74) shows that Φ ′ X (1) gives the mean of X. Using both (30.75) <strong>and</strong>(30.51) allows us to writeΦ ′′ X(1) + Φ ′ X(1) − [ Φ ′ X(1) ] 2= E[X(X − 1)] + E[X] − (E[X])2= E [ X 2] − E[X]+E[X] − (E[X]) 2= E [ X 2] − (E[X]) 2= V [X], (30.76)<strong>and</strong> so express the variance of X in terms of the derivatives of its probabilitygenerating function.◮A r<strong>and</strong>om variable X is given by the number of trials needed to obtain a first successwhen the chance of success at each trial is constant <strong>and</strong> equal to p. Find the probabilitygenerating function <strong>for</strong> X <strong>and</strong> use it to determine the mean <strong>and</strong> variance of X.Clearly, at least one trial is needed, <strong>and</strong> so f 0 =0.Ifn (≥ 1) trials are needed <strong>for</strong> the firstsuccess, the first n − 1 trials must have resulted in failure. ThusPr(X = n) =q n−1 p, n ≥ 1, (30.77)where q =1− p is the probability of failure in each individual trial.The corresponding probability generating function is thus∞∑∞∑Φ X (t) = f n t n = (q n−1 p)t nn=0n=1= p ∞∑(qt) n = p qq × qt1 − qt = pt1 − qt , (30.78)n=1where we have used the result <strong>for</strong> the sum of a geometric series, given in chapter 4, toobtain a closed-<strong>for</strong>m expression <strong>for</strong> Φ X (t). Again, as must be the case, Φ X (1) = 1.To find the mean <strong>and</strong> variance of X we need to evaluate Φ ′ X (1) <strong>and</strong> Φ′′ X (1). Differentiating(30.78) givesΦ ′ pX(t) =⇒ Φ ′(1 − qt)X(1) = p 2 p = 1 2 p ,Φ ′′ X(t) =2pq ⇒ Φ ′′(1 − qt)X(1) = 2pq = 2q3 p 3 p . 2Thus, using (30.74) <strong>and</strong> (30.76),E[X] =Φ ′ X(1) = 1 p ,V [X] =Φ ′′ X(1) + Φ ′ X(1) − [Φ ′ X(1)] 2= 2qp 2 + 1 p − 1 p 2 = q p 2 .A distribution with probabilities of the general <strong>for</strong>m (30.77) is known as a geometricdistribution <strong>and</strong> is discussed in subsection 30.8.2. This <strong>for</strong>m of distribution is common in‘waiting time’ problems (subsection 30.9.3). ◭1159


PROBABILITYnr = nrFigure 30.10The pairs of values of n <strong>and</strong> r used in the evaluation of Φ X+Y (t).Sums of r<strong>and</strong>om variablesWe now turn to considering the sum of two or more independent r<strong>and</strong>omvariables, say X <strong>and</strong> Y , <strong>and</strong> denote by S 2 the r<strong>and</strong>om variableS 2 = X + Y.If Φ S2 (t) isthePGF<strong>for</strong>S 2 , the coefficient of t n in its expansion is given by theprobability that X + Y = n <strong>and</strong> is thus equal to the sum of the probabilities thatX = r <strong>and</strong> Y = n − r <strong>for</strong> all values of r in 0 ≤ r ≤ n. Since such outcomes <strong>for</strong>different values of r are mutually exclusive, we havePr(X + Y = n) =∞∑Pr(X = r)Pr(Y = n − r). (30.79)r=0Multiplying both sides of (30.79) by t n <strong>and</strong> summing over all values of n enablesus to express this relationship in terms of probability generating functions asfollows:∞∑∞∑ n∑Φ X+Y (t) = Pr(X + Y = n)t n = Pr(X = r)t r Pr(Y = n − r)t n−rn=0=n=0 r=0∞∑r=0 n=r∞∑Pr(X = r)t r Pr(Y = n − r)t n−r .The change in summation order is justified by reference to figure 30.10, whichillustrates that the summations are over exactly the same pairs of values of n <strong>and</strong>r, but with the first (inner) summation over the points in a column rather thanover the points in a row. Now, setting n = r + s gives the final result,Φ X+Y (t) =∞∑Pr(X = r)t rr=0∞ ∑s=0Pr(Y = s)t s=Φ X (t)Φ Y (t), (30.80)1160


30.7 GENERATING FUNCTIONSi.e. the PGF of the sum of two independent r<strong>and</strong>om variables is equal to theproduct of their individual PGFs. The same result can be deduced in a less <strong>for</strong>malway by noting that if X <strong>and</strong> Y are independent thenE [ t X+Y ] = E [ t X] E [ t Y ] .Clearly result (30.80) can be extended to more than two r<strong>and</strong>om variables bywriting S 3 = S 2 + Z etc., to given∏Φ ∑(n Xi)(t) = Φ Xi (t), (30.81)i=1<strong>and</strong>, further, if all the X i have the same probability distribution,i=1Φ (∑ ni=1 Xi)(t) =[Φ X(t)] n . (30.82)This latter result has immediate application in the deduction of the PGF <strong>for</strong> thebinomial distribution from that <strong>for</strong> a single trial, equation (30.72).Variable-length sums of r<strong>and</strong>om variablesAs a final result in the theory of probability generating functions we show how tocalculate the PGF <strong>for</strong> a sum of N r<strong>and</strong>om variables, all with the same probabilitydistribution, when the value of N is itself a r<strong>and</strong>om variable but one with aknown probability distribution. In symbols, we wish to find the distributionofS N = X 1 + X 2 + ···+ X N , (30.83)where N is a r<strong>and</strong>om variable with Pr(N = n) = h n <strong>and</strong> PGF χ N (t) =∑hn t n .The probability ξ k that S N = k is given by a sum of conditional probabilities,namely § ∞∑ξ k = Pr(N = n)Pr(X 0 + X 1 + X 2 + ···+ X n = k)n=0∞∑= h n × coefficient of t k in [Φ X (t)] n .n=0Multiplying both sides of this equation by t k <strong>and</strong> summing over all k, weobtain§ Formally X 0 = 0 has to be included, since Pr(N = 0) may be non-zero.1161


PROBABILITYan expression <strong>for</strong> the PGF Ξ S (t) ofS N :∞∑∞∑ ∑∞Ξ S (t) = ξ k t k = t k h n × coefficient of t k in [Φ X (t)] nk=0==k=0∞∑n=0h nn=0∑∞k=0∞∑h n [Φ X (t)] nn=0t k × coefficient of t k in [Φ X (t)] n= χ N (Φ X (t)). (30.84)In words, the PGF of the sum S N is given by the compound function χ N (Φ X (t))obtained by substituting Φ X (t) <strong>for</strong>t in the PGF <strong>for</strong> the number of terms N inthe sum. We illustrate this with the following example.◮The probability distribution <strong>for</strong> the number of eggs in a clutch is Poisson distributed withmean λ, <strong>and</strong> the probability that each egg will hatch is p (<strong>and</strong> is independent of the size ofthe clutch). Use the results stated in (30.72) <strong>and</strong> (30.73) to show that the PGF (<strong>and</strong> hencethe probability distribution) <strong>for</strong> the number of chicks that hatch corresponds to a Poissondistribution having mean λp.The number of chicks that hatch is given by a sum of the <strong>for</strong>m (30.83) in which X i =1ifthe ith chick hatches <strong>and</strong> X i = 0 if it does not. As given by (30.72), Φ X (t) is thus (1−p)+pt.The value of N is given by a Poisson distribution with mean λ; thus, from (30.73), in theterminology of our previous discussion,χ N (t) =e −λ e λt .We now substitute these <strong>for</strong>ms into (30.84) to obtainΞ S (t) =exp(−λ)exp[λΦ X (t)]=exp(−λ)exp{λ[(1 − p)+pt]}=exp(−λp)exp(λpt).But this is exactly the PGF of a Poisson distribution with mean λp.That this implies that the probability is Poisson distributed is intuitively obvious since,in the expansion of the PGF as a power series in t, every coefficient will be preciselythat implied by such a distribution. A solution of the same problem by direct calculationappears in the answer to exercise 30.29. ◭30.7.2 Moment generating functionsAs we saw in section 30.5 a probability function is often expressed in terms ofits moments. This leads naturally to the second type of generating function, amoment generating function. For a r<strong>and</strong>om variable X, <strong>and</strong> a real number t, themoment generating function (MGF) is defined byM X (t) =E [ {∑e tX] if(x=etxi i ) <strong>for</strong> a discrete distribution,∫e tx f(x) dx <strong>for</strong> a continuous distribution.(30.85)1162


30.7 GENERATING FUNCTIONSThe MGF will exist <strong>for</strong> all values of t provided that X is bounded <strong>and</strong> alwaysexists at the point t =0whereM(0) = E(1) = 1.It will be apparent that the PGF <strong>and</strong> the MGF <strong>for</strong> a r<strong>and</strong>om variable Xare closely related. The <strong>for</strong>mer is the expectation of t X whilst the latter is theexpectation of e tX :Φ X (t) =E [ t X] , M X (t) =E [ e tX] .The MGF can thus be obtained from the PGF by replacing t by e t , <strong>and</strong> viceversa. The MGF has more general applicability, however, since it can be usedwith both continuous <strong>and</strong> discrete distributions whilst the PGF is restricted tonon-negative integer distributions.As its name suggests, the MGF is particularly useful <strong>for</strong> obtaining the momentsof a distribution, as is easily seen by noting thatE [ e tX] = E[1+tX + t2 X 2 ]+ ···2!=1+E[X]t + E [ X 2] t 22! + ··· .Assuming that the MGF exists <strong>for</strong> all t around the point t = 0, we can deducethat the moments of a distribution are given in terms of its MGF by∣E[X n ]= dn M X (t) ∣∣∣t=0dt n . (30.86)Similarly, by substitution in (30.51), the variance of the distribution is given byV [X] =M ′′ X(0) − [ M ′ X(0) ] 2, (30.87)where the prime denotes differentiation with respect to t.◮The MGF <strong>for</strong> the Gaussian distribution (see the end of subsection 30.9.1) is given byM X (t) =exp ( µt + 1 2 σ2 t 2) .Find the expectation <strong>and</strong> variance of this distribution.Using (30.86),M X(t) ′ = ( µ + σ 2 t ) exp ( µt + 1 2 σ2 t 2) ⇒ E[X] =M X(0) ′ = µ,M X(t) ′′ = [ σ 2 +(µ + σ 2 t) 2] exp ( µt + 1 2 σ2 t 2) ⇒ M X(0) ′′ = σ 2 + µ 2 .Thus, using (30.87),V [X] =σ 2 + µ 2 − µ 2 = σ 2 .That the mean is found to be µ <strong>and</strong> the variance σ 2 justifies the use of these symbols inthe Gaussian distribution. ◭The moment generating function has several useful properties that follow fromits definition <strong>and</strong> can be employed in simplifying calculations.1163


PROBABILITYScaling <strong>and</strong> shiftingIf Y = aX + b, wherea <strong>and</strong> b are arbitrary constants, thenM Y (t) =E [ e tY ] = E [ e t(aX+b)] = e bt E [ e atX] = e bt M X (at). (30.88)This result is often useful <strong>for</strong> obtaining the central moments of a distribution. If theMFG of X is M X (t) then the variable Y = X−µ has the MGF M Y (t) =e −µt M X (t),which clearly generates the central moments of X, i.e.( )E[(X − µ) n ]=E[Y n ]=M (n) dnY (0) = dt n [e−µt M X (t)] .t=0Sums of r<strong>and</strong>om variablesIf X 1 ,X 2 ,...,X N are independent r<strong>and</strong>om variables <strong>and</strong> S N = X 1 + X 2 + ···+ X NthenM SN (t) =E [ e ] tSN = E [ [ N]e t(X1+X2+···+XN)] ∏= E e tXi .Since the X i are independent,N∏M SN (t) = E [ e tXi] N∏= M Xi (t). (30.89)i=1In words, the MGF of the sum of N independent r<strong>and</strong>om variables is the productof their individual MGFs. By combining (30.89) with (30.88), we obtain the moregeneral result that the MGF of S N = c 1 X 1 + c 2 X 2 + ···+ c N X N (where the c i areconstants) is given byM SN (t) =i=1i=1N∏M Xi (c i t). (30.90)i=1Variable-length sums of r<strong>and</strong>om variablesLet us consider the sum of N independent r<strong>and</strong>om variables X i (i =1, 2,...,N), allwith the same probability distribution, <strong>and</strong> let us suppose that N is itself a r<strong>and</strong>omvariable with a known distribution. Following the notation of section 30.7.1,S N = X 1 + X 2 + ···+ X N ,where N is a r<strong>and</strong>om variable with Pr(N = n) =h n <strong>and</strong> probability generatingfunction χ N (t) = ∑ h n t n . For definiteness, let us assume that the X i are continuousRVs (an analogous discussion can be given in the discrete case). Thus, the1164


30.7 GENERATING FUNCTIONSprobability that value of S N lies in the interval s to s + ds is given by §Pr(s


PROBABILITYso that C X (t) =M X (it), where M X (t) istheMGFofX. Clearly, the characteristicfunction <strong>and</strong> the MGF are very closely related <strong>and</strong> can be used interchangeably.Because of the <strong>for</strong>mal similarity between the definitions of C X (t) <strong>and</strong>M X (t), thecharacteristic function possesses analogous properties to those listed in the previoussection <strong>for</strong> the MGF, with only minor modifications. Indeed, by substituting it<strong>for</strong> t in any of the relations obeyed by the MGF <strong>and</strong> noting that C X (t) =M X (it),we obtain the corresponding relationship <strong>for</strong> the characteristic function. Thus, <strong>for</strong>example, the moments of X are given in terms of the derivatives of C X (t) byE[X n ]=(−i) n C (n)X (0).Similarly, if Y = aX + b then C Y (t) =e ibt C X (at).Whether to describe a r<strong>and</strong>om variable by its characteristic function or by itsMGF is partly a matter of personal preference. However, the use of the CF doeshave some advantages. Most importantly, the replacement of the exponential e tXin the definition of the MGF by the complex oscillatory function e itX in the CFmeans that in the latter we avoid any difficulties associated with convergence ofthe relevant sum or integral. Furthermore, when X is a continous RV, we seefrom (30.91) that C X (t) is related to the Fourier trans<strong>for</strong>m of the PDF f(x). Asa consequence of Fourier’s inversion theorem, we may obtain f(x) fromC X (t) byper<strong>for</strong>ming the inverse trans<strong>for</strong>mf(x) = 1 ∫ ∞C X (t)e −itx dt.2π−∞30.7.4 Cumulant generating functionAs mentioned at the end of subsection 30.5.5, we may also describe a probabilitydensity function f(x) in terms of its cumulants. These quantities may be expressedin terms of the moments of the distribution <strong>and</strong> are important in sampling theory,which we discuss in the next chapter. The cumulants of a distribution are bestdefined in terms of its cumulant generating function (CGF), given by K X (t) =ln M X (t) whereM X (t) is the MGF of the distribution. If K X (t) isexp<strong>and</strong>edasapower series in t then the kth cumulant κ k of f(x) is the coefficient of t k /k!:t 2K X (t) =lnM X (t) ≡ κ 1 t + κ 22! + κ t 33 + ··· . (30.92)3!Since M X (0) = 1, K X (t) contains no constant term.◮Find all the cumulants of the Gaussian distribution discussed in the previous example.The moment generating function <strong>for</strong> the Gaussian distribution is M X (t) =exp ( µt + 1 2 σ2 t 2) .Thus, the cumulant generating function has the simple <strong>for</strong>mK X (t) =lnM X (t) =µt + 1 2 σ2 t 2 .1166


30.7 GENERATING FUNCTIONSComparing this expression with (30.92), we find that κ 1 = µ, κ 2 = σ 2 <strong>and</strong> all othercumulants are equal to zero. ◭We may obtain expressions <strong>for</strong> the cumulants of a distribution in terms of itsmoments by differentiating (30.92) with respect to t to givedK X= 1 dM X.dt M X dtExp<strong>and</strong>ing each term as power series in t <strong>and</strong> cross-multiplying, we obtaint(κ 2···)(1 + κ 2 t + κ 32! + t 2···)1+µ 1 t + µ 22! + t=(µ 2···)1 + µ 2 t + µ 32! + ,<strong>and</strong>, on equating coefficients of like powers of t on each side, we findµ 1 = κ 1 ,µ 2 = κ 2 + κ 1 µ 1 ,µ 3 = κ 3 +2κ 2 µ 1 + κ 1 µ 2 ,µ 4 = κ 4 +3κ 3 µ 1 +3κ 2 µ 2 + κ 1 µ 3 ,.µ k = κ k + k−1 C 1 κ k−1 µ 1 + ···+ k−1 C r κ k−r µ r + ···+ κ 1 µ k−1 .Solving these equations <strong>for</strong> the κ k , we obtain (<strong>for</strong> the first four cumulants)κ 1 = µ 1 ,κ 2 = µ 2 − µ 2 1 = ν 2 ,κ 3 = µ 3 − 3µ 2 µ 1 +2µ 3 1 = ν 3 ,κ 4 = µ 4 − 4µ 3 µ 1 +12µ 2 µ 2 1 − 3µ2 2 − 6µ4 1 = ν 4 − 3ν2 2 . (30.93)Higher-order cumulants may be calculated in the same way but become increasinglylengthy to write out in full.The principal property of cumulants is their additivity, which may be provedby combining (30.92) with (30.90). If X 1 , X 2 , ..., X N are independent r<strong>and</strong>omvariables <strong>and</strong> K Xi (t) <strong>for</strong> i =1, 2,...,N is the CGF <strong>for</strong> X i then the CGF ofS N = c 1 X 1 + c 2 X 2 + ···+ c N X N (where the c i are constants) is given byK SN (t) =N∑K Xi (c i t).i=1Cumulants also have the useful property that, under a change of origin X →X + a the first cumulant undergoes the change κ 1 → κ 1 + a but all higher-ordercumulants remain unchanged. Under a change of scale X → bX, cumulant κ rundergoes the change κ r → b r κ r .1167


PROBABILITYDistribution Probability law f(x) MGF E[X] V [X]binomial n C x p x q n−x (pe t + q) n np npq( ) r pnegative binomial r+x−1 C x p r q x rq1 − qe t pgeometric q x−1 pe t 1p1 − qe t phypergeometricPoisson(Np)!(Nq)!n!(N−n)!x!(Np−x)!(n−x)!(Nq−n+x)!N!λ xx! e−λ e λ(et −1)npλrqp 2qp 2N − nN − 1 npqλTable 30.1Some important discrete probability distributions.30.8 Important discrete distributionsHaving discussed some general properties of distributions, we now consider themore important discrete distributions encountered in physical applications. Theseare discussed in detail below, <strong>and</strong> summarised <strong>for</strong> convenience in table 30.1; werefer the reader to the relevant section below <strong>for</strong> an explanation of the symbolsused.30.8.1 The binomial distributionPerhaps the most important discrete probability distribution is the binomial distribution.This distribution describes processes that consist of a number of independentidentical trials with two possible outcomes, A <strong>and</strong> B = Ā. We may callthese outcomes ‘success’ <strong>and</strong> ‘failure’ respectively. If the probability of a successis Pr(A) =p then the probability of a failure is Pr(B) =q =1− p. Ifweper<strong>for</strong>mn trials then the discrete r<strong>and</strong>om variableX = number of times A occurscan take the values 0, 1, 2,...,n; its distribution amongst these values is describedby the binomial distribution.We now calculate the probability that in n trials we obtain x successes (<strong>and</strong> son − x failures). One way of obtaining such a result is to have x successes followedby n−x failures. Since the trials are assumed independent, the probability of this ispp ···p ×} {{ }qq ···q} {{ }= p x q n−x .x times n − x timesThis is, however, just one permutation of x successes <strong>and</strong> n − x failures. The total1168


30.8 IMPORTANT DISCRETE DISTRIBUTIONSf(x)f(x)n =5,p =0.6 n =5,p =0.1670.40.40.30.30.20.20.1 0.100 1 2 3 4 5 x00 1 2 3 4 5 xf(x)f(x)n = 10, p =0.6 n = 10, p =0.1670.4 0.40.3 0.30.2 0.20.10.10 0 1 2 3 4 5 6 7 8 9 10 x0 0 1 2 3 4 5 6 7 8 910 xFigure 30.11 Some typical binomial distributions with various combinationsof parameters n <strong>and</strong> p.number of permutations of n objects, of which x are identical <strong>and</strong> of type 1 <strong>and</strong>n − x are identical <strong>and</strong> of type 2, is given by (30.33) asn!x!(n − x)! ≡ n C x .There<strong>for</strong>e, the total probability of obtaining x successes from n trials isf(x) =Pr(X = x) = n C x p x q n−x = n C x p x (1 − p) n−x , (30.94)which is the binomial probability distribution <strong>for</strong>mula. When a r<strong>and</strong>om variableX follows the binomial distribution <strong>for</strong> n trials, with a probability of success p,we write X ∼ Bin(n, p). Then the r<strong>and</strong>om variable X is often referred to as abinomial variate. Some typical binomial distributions are shown in figure 30.11.◮If a single six-sided die is rolled five times, what is the probability that a six is thrownexactly three times?Here the number of ‘trials’ n = 5, <strong>and</strong> we are interested in the r<strong>and</strong>om variableX = number of sixes thrown.Since the probability of a ‘success’ is p = 1 , the probability of obtaining exactly three sixes6in five throws is given by (30.94) as( ) 3 ( ) (5−3)5! 1 5Pr(X =3)==0.032. ◭3!(5 − 3)! 6 61169


PROBABILITYFor evaluating binomial probabilities a useful result is the binomial recurrence<strong>for</strong>mulaPr(X = x +1)= p ( ) n − xPr(X = x), (30.95)q x +1which enables successive probabilities Pr(X = x + k), k =1, 2,..., to be calculatedonce Pr(X = x) is known; it is often quicker to use than (30.94).◮The r<strong>and</strong>om variable X is distributed as X ∼ Bin(3, 1 ). Evaluate the probability function2f(x) using the binomial recurrence <strong>for</strong>mula.The probability Pr(X = 0) may be calculated using (30.94) <strong>and</strong> isThe ratio p/q = 1 / 1 2 2(30.95), we findPr(X =0)= 3 C 0( 12) 0 ( 12) 3=18 .= 1 in this case <strong>and</strong> so, using the binomial recurrence <strong>for</strong>mulaPr(X =1)=1× 3 − 00+1 × 1 8 = 3 8 ,Pr(X =2)=1× 3 − 11+1 × 3 8 = 3 8 ,Pr(X =3)=1× 3 − 22+1 × 3 8 = 1 8 ,results which may be verified by direct application of (30.94). ◭We note that, as required, the binomial distribution satifiesn∑f(x) =x=0n∑n C x p x q n−x =(p + q) n =1.x=0Furthermore, from the definitions of E[X] <strong>and</strong>V [X] <strong>for</strong> a discrete distribution,we may show that <strong>for</strong> the binomial distribution E[X] =np <strong>and</strong> V [X] =npq. Thedirect summations involved are, however, rather cumbersome <strong>and</strong> these resultsare obtained much more simply using the moment generating function.The moment generating function <strong>for</strong> the binomial distributionTo find the MGF <strong>for</strong> the binomial distribution we consider the binomial r<strong>and</strong>omvariable X to be the sum of the r<strong>and</strong>om variables X i , i =1, 2,...,n, which aredefined by{1 if a ‘success’ occurs on the ith trial,X i =0 if a ‘failure’ occurs on the ith trial.1170


30.8 IMPORTANT DISCRETE DISTRIBUTIONSThusM i (t) =E [ e tXi] = e 0t × Pr(X i =0)+e 1t × Pr(X i =1)=1× q + e t × p= pe t + q.From (30.89), it follows that the MGF <strong>for</strong> the binomial distribution is given byn∏M(t) = M i (t) =(pe t + q) n . (30.96)i=1We can now use the moment generating function to derive the mean <strong>and</strong>variance of the binomial distribution. From (30.96)M ′ (t) =npe t (pe t + q) n−1 ,<strong>and</strong> from (30.86)E[X] =M ′ (0) = np(p + q) n−1 = np,where the last equality follows from p + q =1.Differentiating with respect to t once more givesM ′′ (t) =e t (n − 1)np 2 (pe t + q) n−2 + e t np(pe t + q) n−1 ,<strong>and</strong> from (30.86)E[X 2 ]=M ′′ (0) = n 2 p 2 − np 2 + np.Thus, using (30.87)V [X] =M ′′ (0) − [ M ′ (0) ] 2= n 2 p 2 − np 2 + np − n 2 p 2 = np(1 − p) =npq.Multiple binomial distributionsSuppose X <strong>and</strong> Y are two independent r<strong>and</strong>om variables, both of which aredescribed by binomial distributions with a common probability of success p, butwith (in general) different numbers of trials n 1 <strong>and</strong> n 2 ,sothatX ∼ Bin(n 1 ,p)<strong>and</strong> Y ∼ Bin(n 2 ,p). Now consider the r<strong>and</strong>om variable Z = X + Y .Wecouldcalculate the probability distribution of Z directly using (30.60), but it is mucheasier to use the MGF (30.96).Since X <strong>and</strong> Y are independent r<strong>and</strong>om variables, the MGF M Z (t) of the newvariable Z = X + Y is given simply by the product of the individual MGFsM X (t) <strong>and</strong>M Y (t). Thus, we obtainM Z (t) =M X (t)M Y (t) =(pe t + q) n1 (pe t + q) n1 =(pe t + q) n1+n2 ,which we recognise as the MGF of Z ∼ Bin(n 1 + n 2 ,p). Hence Z is also describedby a binomial distribution.This result may be extended to any number of binomial distributions. If X i ,1171


PROBABILITYi =1, 2,...,N, is distributed as X i ∼ Bin(n i ,p)thenZ = X 1 + X 2 + ···+ X N isdistributed as Z ∼ Bin(n 1 + n 2 + ···+ n N ,p), as would be expected since the resultof ∑ i n i trials cannot depend on how they are split up. A similar proof is alsopossible using either the probability or cumulant generating functions.Un<strong>for</strong>tunately, no equivalent simple result exists <strong>for</strong> the probability distributionof the difference Z = X − Y of two binomially distributed variables.30.8.2 The geometric <strong>and</strong> negative binomial distributionsA special case of the binomial distribution occurs when instead of the number ofsuccesses we consider the discrete r<strong>and</strong>om variableX = number of trials required to obtain the first success.The probability that x trials are required in order to obtain the first success, issimply the probability of obtaining x − 1 failures followed by one success. If theprobability of a success on each trial is p, then<strong>for</strong>x>0f(x) =Pr(X = x) =(1− p) x−1 p = q x−1 p,where q =1− p. This distribution is sometimes called the geometric distribution.The probability generating function <strong>for</strong> this distribution is given in (30.78). Byreplacing t by e t in (30.78) we immediately obtain the MGF of the geometricdistributionM(t) =pet1 − qe t ,from which its mean <strong>and</strong> variance are found to beE[X] = 1 p , V[X] = q p 2 .Another distribution closely related to the binomial is the negative binomialdistribution. This describes the probability distribution of the r<strong>and</strong>om variableX = number of failures be<strong>for</strong>e the rth success.One way of obtaining x failures be<strong>for</strong>e the rth success is to have r − 1 successesfollowed by x failures followed by the rth success, <strong>for</strong> which the probability ispp ···p} {{ }r − 1times× qq ···q × p = p r q x .} {{ }x timesHowever, the first r + x − 1 factors constitute just one permutation of r − 1successes <strong>and</strong> x failures. The total number of permutations of these r + x − 1objects, of which r − 1 are identical <strong>and</strong> of type 1 <strong>and</strong> x are identical <strong>and</strong> of type1172


30.8 IMPORTANT DISCRETE DISTRIBUTIONS2, is r+x−1 C x . There<strong>for</strong>e, the total probability of obtaining x failures be<strong>for</strong>e therth success isf(x) =Pr(X = x) = r+x−1 C x p r q x ,which is called the negative binomial distribution (see the related discussion onp. 1137). It is straight<strong>for</strong>ward to show that the MGF of this distribution is( ) p rM(t) =1 − qe t ,<strong>and</strong> that its mean <strong>and</strong> variance are given byE[X] = rq <strong>and</strong> V [X] = rqpp 2 .30.8.3 The hypergeometric distributionIn subsection 30.8.1 we saw that the probability of obtaining x successes in nindependent trials was given by the binomial distribution. Suppose that these n‘trials’ actually consist of drawing at r<strong>and</strong>om n balls, from a set of N such ballsof which M are red <strong>and</strong> the rest white. Let us consider the r<strong>and</strong>om variableX = number of red balls drawn.On the one h<strong>and</strong>, if the balls are drawn with replacement then the trials areindependent <strong>and</strong> the probability of drawing a red ball is p = M/N each time.There<strong>for</strong>e, the probability of drawing x red balls in n trials is given by thebinomial distribution asn!Pr(X = x) =x!(n − x)! px (1 − p) n−x .On the other h<strong>and</strong>, if the balls are drawn without replacement the trials are notindependent <strong>and</strong> the probability of drawing a red ball depends on how many redballs have already been drawn. We can, however, still derive a general <strong>for</strong>mula<strong>for</strong> the probability of drawing x red balls in n trials, as follows.The number of ways of drawing x red balls from M is M C x , <strong>and</strong> the numberof ways of drawing n − x white balls from N − M is N−M C n−x . There<strong>for</strong>e, thetotal number of ways to obtain x red balls in n trials is M C N−M x C n−x . However,the total number of ways of drawing n objects from N is simply N C n . Hence theprobability of obtaining x red balls in n trials isM C N−M x C n−xPr(X = x) =NC nM!(N − M)!=x!(M − x)! (n − x)!(N − M − n + x)!=n!(N − n)!, (30.97)N!(Np)!(Nq)! n!(N − n)!x!(Np − x)!(n − x)!(Nq − n + x)! N! , (30.98)1173


PROBABILITYwhere in the last line p = M/N <strong>and</strong> q =1− p. This is called the hypergeometricdistribution.By per<strong>for</strong>ming the relevant summations directly, it may be shown that thehypergeometric distribution has meanE[X] =n M N = np<strong>and</strong> varianceV [X] =nM(N − M)(N − n)N 2 (N − 1)= N − nN − 1 npq.◮In the UK National Lottery each participant chooses six different numbers between 1<strong>and</strong> 49. In each weekly draw six numbered winning balls are subsequently drawn. Find theprobabilities that a participant chooses 0, 1, 2, 3, 4, 5, 6 winning numbers correctly.The probabilities are given by a hypergeometric distribution with N (the total number ofballs) = 49, M (the number of winning balls drawn) = 6, <strong>and</strong> n (the number of numberschosen by each participant) = 6. Thus, substituting in (30.97), we find6 C 43 0 C 6Pr(0) = = 16 49C 6 2.29 , Pr(1) = C 43 1 C 5= 149C 6 2.42 ,6 C 43 2 C 4Pr(2) = = 16 49C 6 7.55 , Pr(3) = C 43 3 C 3= 149C 6 56.6 ,6 C 43 4 C 2Pr(4) = = 16 49C 6 1032 , Pr(5) = C 43 5 C 1= 149C 6 54 200 ,It can easily be seen thatPr(6) =6 C 43 6 C 0 1=49C 6 13.98 × 10 . 6as expected. ◭6∑Pr(i) =0.44 + 0.41 + 0.13 + 0.02 + O(10 −3 )=1,i=0Note that if the number of trials (balls drawn) is small compared with N, M<strong>and</strong> N − M then not replacing the balls is of little consequence, <strong>and</strong> we mayapproximate the hypergeometric distribution by the binomial distribution (withp = M/N); this is much easier to evaluate.30.8.4 The Poisson distributionWe have seen that the binomial distribution describes the number of successfuloutcomes in a certain number of trials n. The Poisson distribution also describesthe probability of obtaining a given number of successes but <strong>for</strong> situationsin which the number of ‘trials’ cannot be enumerated; rather it describes thesituation in which discrete events occur in a continuum. Typical examples of1174


30.8 IMPORTANT DISCRETE DISTRIBUTIONSdiscrete r<strong>and</strong>om variables X described by a Poisson distribution are the numberof telephone calls received by a switchboard in a given interval, or the numberof stars above a certain brightness in a particular area of the sky. Given a meanrate of occurrence λ of these events in the relevant interval or area, the Poissondistribution gives the probability Pr(X = x) that exactly x events will occur.We may derive the <strong>for</strong>m of the Poisson distribution as the limit of the binomialdistribution when the number of trials n →∞<strong>and</strong> the probability of ‘success’p → 0, in such a way that np = λ remains finite. Thus, in our example of atelephone switchboard, suppose we wish to find the probability that exactly xcalls are received during some time interval, given that the mean number of callsin such an interval is λ. Let us begin by dividing the time interval into a largenumber, n, of equal shorter intervals, in each of which the probability of receivinga call is p. Asweletn →∞then p → 0, but since we require the mean numberof calls in the interval to equal λ, we must have np = λ. The probability of xsuccesses in n trials is given by the binomial <strong>for</strong>mula asn!Pr(X = x) =x!(n − x)! px (1 − p) n−x . (30.99)Now as n →∞, with x finite, the ratio of the n-dependent factorials in (30.99)behaves asymptotically as a power of n, i.e.n!limn→∞ (n − x)! = lim n(n − 1)(n − 2) ···(n − x +1)∼ n→∞ nx .Alsolim lim(1− (1 − p) λ/pn→∞ p→0 p)n−x = limp→0 (1 − p) x = e−λ1 .Thus, using λ = np, (30.99) tends to the Poisson distributionf(x) =Pr(X = x) = e−λ λ x, (30.100)x!which gives the probability of obtaining exactly x calls in the given time interval.As we shall show below, λ is the mean of the distribution. Events following aPoisson distribution are usually said to occur r<strong>and</strong>omly in time.Alternatively we may derive the Poisson distribution directly, without consideringa limit of the binomial distribution. Let us again consider our exampleof a telephone switchboard. Suppose that the probability that x calls have beenreceived in a time interval t is P x (t). If the average number of calls received in aunit time is λ then in a further small time interval ∆t the probability of receivinga call is λ∆t, provided ∆t is short enough that the probability of receiving two ormore calls in this small interval is negligible. Similarly the probability of receivingno call during the same small interval is simply 1 − λ∆t.Thus, <strong>for</strong> x>0, the probability of receiving exactly x calls in the total interval1175


PROBABILITYt +∆t is given byP x (t +∆t) =P x (t)(1 − λ∆t)+P x−1 (t)λ∆t.Rearranging the equation, dividing through by ∆t <strong>and</strong> letting ∆t → 0, we obtainthe differential recurrence equationdP x (t)= λP x−1 (t) − λP x (t). (30.101)dtFor x = 0 (i.e. no calls received), however, (30.101) simplifies todP 0 (t)= −λP 0 (t),dtwhich may be integrated to give P 0 (t) =P 0 (0)e −λt . But since the probability P 0 (0)of receiving no calls in a zero time interval must equal unity, we have P 0 (t) =e −λt .This expression <strong>for</strong> P 0 (t) may then be substituted back into (30.101) with x =1to obtain a differential equation <strong>for</strong> P 1 (t) that has the solution P 1 (t) =λte −λt .We may repeat this process to obtain expressions <strong>for</strong> P 2 (t),P 3 (t),...,P x (t), <strong>and</strong> wefindP x (t) = (λt)xx! e−λt . (30.102)By setting t = 1 in (30.102), we again obtain the Poisson distribution (30.100) <strong>for</strong>obtaining exactly x calls in a unit time interval.If a discrete r<strong>and</strong>om variable is described by a Poisson distribution of mean λthen we write X ∼ Po(λ). As it must be, the sum of the probabilities is unity:∞∑Pr(X = x) =e −λx=0∞ ∑x=0λ xx! = e−λ e λ =1.From (30.100) we may also derive the Poisson recurrence <strong>for</strong>mula,Pr(X = x +1)=λx +1 Pr(X = x) <strong>for</strong> x =0, 1, 2,..., (30.103)which enables successive probabilities to be calculated easily once one is known.◮A person receives on average one e-mail message per half-hour interval. Assuming thatthe e-mails are received r<strong>and</strong>omly in time, find the probabilities that in any particular hour0, 1, 2, 3, 4, 5 messages are received.Let X = number of e-mails received per hour. Clearly the mean number of e-mails perhour is two, <strong>and</strong> so X follows a Poisson distribution with λ =2,i.e.Pr(X = x) = 2xx! e−2 .Thus Pr(X =0)=e −2 =0.135, Pr(X =1)=2e −2 =0.271, Pr(X =2)=2 2 e −2 /2! = 0.271,Pr(X =3)=2 3 e −2 /3! = 0.180, Pr(X =4)=2 4 e −2 /4! = 0.090, Pr(X =5)=2 5 e −2 /5! =0.036. These results may also be calculated using the recurrence <strong>for</strong>mula (30.103). ◭1176


30.8 IMPORTANT DISCRETE DISTRIBUTIONS0.3f(x)f(x)λ =1 λ =20.30.20.20.1 0.10001 2 3 4 5 x0 1 2 3 4 5 67xf(x)0.3λ =50.20.1001234 567 8 9 10 11xFigure 30.12 Three Poisson distributions <strong>for</strong> different values of the parameterλ.The above example illustrates the point that a Poisson distribution typicallyrises <strong>and</strong> then falls. It either has a maximum when x is equal to the integer partof λ or, if λ happens to be an integer, has equal maximal values at x = λ − 1<strong>and</strong>x = λ. The Poisson distribution always has a long ‘tail’ towards higher values of Xbut the higher the value of the mean the more symmetric the distribution becomes.Typical Poisson distributions are shown in figure 30.12. Using the definitions ofmean <strong>and</strong> variance, we may show that, <strong>for</strong> the Poisson distribution, E[X] =λ <strong>and</strong>V [X] =λ. Nevertheless, as in the case of the binomial distribution, per<strong>for</strong>mingthe relevant summations directly is rather tiresome, <strong>and</strong> these results are muchmore easily proved using the MGF.The moment generating function <strong>for</strong> the Poisson distributionThe MGF of the Poisson distribution is given byM X (t) =E [ e tX] =∞∑ e tx e −λ λ x ∑∞= e −λx!x=01177x=0(λe t ) xx!= e −λ e λet = e λ(et −1)(30.104)


PROBABILITYfrom which we obtainM X(t) ′ =λe t e λ(et−1) ,M X(t) ′′ =(λ 2 e 2t + λe t )e λ(et−1) .Thus, the mean <strong>and</strong> variance of the Poisson distribution are given byE[X] =M ′ X(0) = λ <strong>and</strong> V [X] =M ′′ X(0) − [M ′ X(0)] 2 = λ.The Poisson approximation to the binomial distributionEarlier we derived the Poisson distribution as the limit of the binomial distributionwhen n →∞<strong>and</strong> p → 0 in such a way that np = λ remains finite, where λ is themean of the Poisson distribution. It is not surprising, there<strong>for</strong>e, that the Poissondistribution is a very good approximation to the binomial distribution <strong>for</strong> largen (≥ 50, say) <strong>and</strong> small p (≤ 0.1, say). Moreover, it is easier to calculate as itinvolves fewer factorials.◮In a large batch of light bulbs, the probability that a bulb is defective is 0.5%. Forasample of 200 bulbs taken at r<strong>and</strong>om, find the approximate probabilities that 0, 1 <strong>and</strong> 2 ofthe bulbs respectively are defective.Let the r<strong>and</strong>om variable X = number of defective bulbs in a sample. This is distributedas X ∼ Bin(200, 0.005), implying that λ = np =1.0. Since n is large <strong>and</strong> p small, we mayapproximate the distribution as X ∼ Po(1), giving−1 1xPr(X = x) ≈ ex! ,from which we find Pr(X =0)≈ 0.37, Pr(X =1)≈ 0.37, Pr(X =2)≈ 0.18. For comparison,it may be noted that the exact values calculated from the binomial distribution are identicalto those found here to two decimal places. ◭Multiple Poisson distributionsMirroring our discussion of multiple binomial distributions in subsection 30.8.1,let us suppose X <strong>and</strong> Y are two independent r<strong>and</strong>om variables, both of whichare described by Poisson distributions with (in general) different means, so thatX ∼ Po(λ 1 )<strong>and</strong>Y ∼ Po(λ 2 ). Now consider the r<strong>and</strong>om variable Z = X + Y .Wemay calculate the probability distribution of Z directly using (30.60), but we mayderive the result much more easily by using the moment generating function (orindeed the probability or cumulant generating functions).Since X <strong>and</strong> Y are independent RVs, the MGF <strong>for</strong> Z is simply the product ofthe individual MGFs <strong>for</strong> X <strong>and</strong> Y . Thus, from (30.104),M Z (t) =M X (t)M Y (t) =e λ1(et −1) e λ2(et −1) = e (λ1+λ2)(et −1) ,which we recognise as the MGF of Z ∼ Po(λ 1 + λ 2 ). Hence Z is also Poissondistributed <strong>and</strong> has mean λ 1 + λ 2 . Un<strong>for</strong>tunately, no such simple result holds <strong>for</strong>the difference Z = X − Y of two independent Poisson variates. A closed-<strong>for</strong>m1178


30.9 IMPORTANT CONTINUOUS DISTRIBUTIONSexpression <strong>for</strong> the PDF of this Z does exist, but it is a rather complicatedcombination of exponentials <strong>and</strong> a modified Bessel function. §◮Two types of e-mail arrive independently <strong>and</strong> at r<strong>and</strong>om: external e-mails at a mean rateof one every five minutes <strong>and</strong> internal e-mails at a rate of two every five minutes. Calculatethe probability of receiving two or more e-mails in any two-minute interval.LetX = number of external e-mails per two-minute interval,Y = number of internal e-mails per two-minute interval.Since we expect on average one external e-mail <strong>and</strong> two internal e-mails every five minuteswe have X ∼ Po(0.4) <strong>and</strong> Y ∼ Po(0.8). Letting Z = X + Y we have Z ∼ Po(0.4+0.8) =Po(1.2). NowPr(Z ≥ 2) = 1 − Pr(Z


PROBABILITYDistribution Probability law f(x) MGF E[X] V [X][ ]1Gaussianσ √ 2π exp (x − µ)2− exp(µt + 12σ 2 2 σ2 t 2 ) µ σ 2( ) λexponential λe −λx 1 1λ − tλ λ( ) 2rλλgammaΓ(r) (λx)r−1 e −λx r rλ − tλ λ 2( ) n/211chi-squared2 n/2 Γ(n/2) x(n/2)−1 e −x/2 n 2n1 − 2t1e bt − e at a + b (b − a) 2uni<strong>for</strong>mb − a(b − a)t2 12Table 30.2Some important continuous probability distributions.The probability density function <strong>for</strong> a Gaussian distribution of a r<strong>and</strong>omvariable X, with mean E[X] =µ <strong>and</strong> variance V [X] =σ 2 ,takesthe<strong>for</strong>mf(x) = 1σ √ 2π exp [− 1 2( x − µ) 2]. (30.105)σThe factor 1/ √ 2π arises from the normalisation of the distribution,∫ ∞−∞f(x)dx =1;the evaluation of this integral is discussed in subsection 6.4.2. The Gaussi<strong>and</strong>istribution is symmetric about the point x = µ <strong>and</strong> has the characteristic ‘bell’shape shown in figure 30.13. The width of the curve is described by the st<strong>and</strong>arddeviation σ: ifσ is large then the curve is broad, <strong>and</strong> if σ is small then the curveis narrow (see the figure). At x = µ ± σ, f(x) falls to e −1/2 ≈ 0.61 of its peakvalue; these points are points of inflection, where d 2 f/dx 2 = 0. When a r<strong>and</strong>omvariable X follows a Gaussian distribution with mean µ <strong>and</strong> variance σ 2 ,wewriteX ∼ N(µ, σ 2 ).The effects of changing µ <strong>and</strong> σ are only to shift the curve along the x-axis orto broaden or narrow it, respectively. Thus all Gaussians are equivalent in thata change of origin <strong>and</strong> scale can reduce them to a st<strong>and</strong>ard <strong>for</strong>m. We there<strong>for</strong>econsider the r<strong>and</strong>om variable Z =(X − µ)/σ, <strong>for</strong> which the PDF takes the <strong>for</strong>mφ(z) = 1 √2πexp(− z22), (30.106)which is called the st<strong>and</strong>ard Gaussian distribution <strong>and</strong> has mean µ = 0 <strong>and</strong>variance σ 2 = 1. The r<strong>and</strong>om variable Z is called the st<strong>and</strong>ard variable.From (30.105) we can define the cumulative probability function <strong>for</strong> a Gaussian1180


30.9 IMPORTANT CONTINUOUS DISTRIBUTIONS0.40.30.2µ =3σ =1σ =20.1σ =3−6 −4 −2 2 3 4 6 8 10 12Figure 30.13 The Gaussian or normal distribution <strong>for</strong> mean µ =3<strong>and</strong>various values of the st<strong>and</strong>ard deviation σ.Φ(a)0.40.30.20.1φ(z)Φ(a)Φ(z)1z−2 0 a−4 a 2 4 −2 −1 y 2Figure 30.14 On the left, the st<strong>and</strong>ard Gaussian distribution φ(z); the shadedarea gives Pr(Z


PROBABILITYIt is usual only to tabulate Φ(z) <strong>for</strong>z>0, since it can be seen easily, fromfigure 30.14 <strong>and</strong> the symmetry of the Gaussian distribution, that Φ(−z) =1−Φ(z);see table 30.3. Using such a table it is then straight<strong>for</strong>ward to evaluate theprobability that Z lies in a given range of z-values. For example, <strong>for</strong> a <strong>and</strong> bconstant,Pr(Z a)=1− Φ(a),Pr(a


30.9 IMPORTANT CONTINUOUS DISTRIBUTIONSΦ(z) .00 .01 .02 .03 .04 .05 .06 .07 .08 .090.0 .5000 .5040 .5080 .5120 .5160 .5199 .5239 .5279 .5319 .53590.1 .5398 .5438 .5478 .5517 .5557 .5596 .5636 .5675 .5714 .57530.2 .5793 .5832 .5871 .5910 .5948 .5987 .6026 .6064 .6103 .61410.3 .6179 .6217 .6255 .6293 .6331 .6368 .6406 .6443 .6480 .65170.4 .6554 .6591 .6628 .6664 .6700 .6736 .6772 .6808 .6844 .68790.5 .6915 .6950 .6985 .7019 .7054 .7088 .7123 .7157 .7190 .72240.6 .7257 .7291 .7324 .7357 .7389 .7422 .7454 .7486 .7517 .75490.7 .7580 .7611 .7642 .7673 .7704 .7734 .7764 .7794 .7823 .78520.8 .7881 .7910 .7939 .7967 .7995 .8023 .8051 .8078 .8106 .81330.9 .8159 .8186 .8212 .8238 .8264 .8289 .8315 .8340 .8365 .83891.0 .8413 .8438 .8461 .8485 .8508 .8531 .8554 .8577 .8599 .86211.1 .8643 .8665 .8686 .8708 .8729 .8749 .8770 .8790 .8810 .88301.2 .8849 .8869 .8888 .8907 .8925 .8944 .8962 .8980 .8997 .90151.3 .9032 .9049 .9066 .9082 .9099 .9115 .9131 .9147 .9162 .91771.4 .9192 .9207 .9222 .9236 .9251 .9265 .9279 .9292 .9306 .93191.5 .9332 .9345 .9357 .9370 .9382 .9394 .9406 .9418 .9429 .94411.6 .9452 .9463 .9474 .9484 .9495 .9505 .9515 .9525 .9535 .95451.7 .9554 .9564 .9573 .9582 .9591 .9599 .9608 .9616 .9625 .96331.8 .9641 .9649 .9656 .9664 .9671 .9678 .9686 .9693 .9699 .97061.9 .9713 .9719 .9726 .9732 .9738 .9744 .9750 .9756 .9761 .97672.0 .9772 .9778 .9783 .9788 .9793 .9798 .9803 .9808 .9812 .98172.1 .9821 .9826 .9830 .9834 .9838 .9842 .9846 .9850 .9854 .98572.2 .9861 .9864 .9868 .9871 .9875 .9878 .9881 .9884 .9887 .98902.3 .9893 .9896 .9898 .9901 .9904 .9906 .9909 .9911 .9913 .99162.4 .9918 .9920 .9922 .9925 .9927 .9929 .9931 .9932 .9934 .99362.5 .9938 .9940 .9941 .9943 .9945 .9946 .9948 .9949 .9951 .99522.6 .9953 .9955 .9956 .9957 .9959 .9960 .9961 .9962 .9963 .99642.7 .9965 .9966 .9967 .9968 .9969 .9970 .9971 .9972 .9973 .99742.8 .9974 .9975 .9976 .9977 .9977 .9978 .9979 .9979 .9980 .99812.9 .9981 .9982 .9982 .9983 .9984 .9984 .9985 .9985 .9986 .99863.0 .9987 .9987 .9987 .9988 .9988 .9989 .9989 .9989 .9990 .99903.1 .9990 .9991 .9991 .9991 .9992 .9992 .9992 .9992 .9993 .99933.2 .9993 .9993 .9994 .9994 .9994 .9994 .9994 .9995 .9995 .99953.3 .9995 .9995 .9995 .9996 .9996 .9996 .9996 .9996 .9996 .99973.4 .9997 .9997 .9997 .9997 .9997 .9997 .9997 .9997 .9997 .9998Table 30.3 The cumulative probability function Φ(z) <strong>for</strong> the st<strong>and</strong>ard Gaussi<strong>and</strong>istribution, as given by (30.108). The units <strong>and</strong> the first decimal placeof z are specified in the column under Φ(z) <strong>and</strong> the second decimal place isspecified by the column headings. Thus, <strong>for</strong> example, Φ(1.23) = 0.8907.1183


PROBABILITY◮Sawmill A produces boards whose lengths are Gaussian distributed with mean 209.4 cm<strong>and</strong> st<strong>and</strong>ard deviation 5.0 cm. A board is accepted if it is longer than 200 cm but isrejected otherwise. Show that 3% of boards are rejected.Sawmill B produces boards of the same st<strong>and</strong>ard deviation but of mean length 210.1 cm.Find the proportion of boards rejected if they are drawn at r<strong>and</strong>om from the outputs of A<strong>and</strong> B in the ratio 3:1.Let X = length of boards from A, sothatX ∼ N(209.4, (5.0) 2 )<strong>and</strong>( ) ( )200 − µ 200 − 209.4Pr(X


30.9 IMPORTANT CONTINUOUS DISTRIBUTIONSNow using Φ(−z) =1− Φ(z) gives( ) µ − 140Φ=1− 0.030 = 0.970.σUsing table 30.3 again, we findµ − 140=1.88. (30.113)σSolving the simultaneous equations (30.112) <strong>and</strong> (30.113) gives µ = 173.5, σ =17.8. ◭The moment generating function <strong>for</strong> the Gaussian distributionUsing the definition of the MGF (30.85),M X (t) =E [ e tX] ∫ ∞[]1=−∞ σ √ 2π exp (x − µ)2tx −2σ 2 dx= c exp ( µt + 1 2 σ2 t 2) ,where the final equality is established by completing the square in the argumentof the exponential <strong>and</strong> writing∫ ∞{1c =−∞ σ √ 2π exp − [x − (µ + σ2 t)] 2 }2σ 2 dx.However, the final integral is simply the normalisation integral <strong>for</strong> the Gaussi<strong>and</strong>istribution, <strong>and</strong> so c = 1 <strong>and</strong> the MGF is given byM X (t) =exp ( µt + 1 2 σ2 t 2) . (30.114)We showed in subsection 30.7.2 that this MGF leads to E[X] =µ <strong>and</strong> V [X] =σ 2 ,as required.Gaussian approximation to the binomial distributionWe may consider the Gaussian distribution as the limit of the binomial distributionwhen the number of trials n →∞but the probability of a success p remainsfinite, so that np →∞also. (This contrasts with the Poisson distribution, whichcorresponds to the limit n →∞<strong>and</strong> p → 0 with np = λ remaining finite.) Inother words, a Gaussian distribution results when an experiment with a finiteprobability of success is repeated a large number of times. We now show howthis Gaussian limit arises.The binomial probability function gives the probability of x successes in n trialsasn!f(x) =x!(n − x)! px (1 − p) n−x .Taking the limit as n →∞(<strong>and</strong> x →∞) we may approximate the factorials byStirling’s approximationn! ∼ √ ( n) n2πne1185


PROBABILITYx f(x) (binomial) f(x) (Gaussian)0 0.0001 0.00011 0.0016 0.00142 0.0106 0.00923 0.0425 0.03954 0.1115 0.11195 0.2007 0.20916 0.2508 0.25757 0.2150 0.20918 0.1209 0.11199 0.0403 0.039510 0.0060 0.0092Table 30.4 Comparison of the binomial distribution <strong>for</strong> n =10<strong>and</strong>p =0.6with its Gaussian approximation.to obtainf(x) ≈ √ 1 ( x) −x−1/2 ( n − x2πn nn= √ 1 [exp − ( )x + 1 x2πn2 ln) −n+x−1/2p x (1 − p) n−xn − ( n − x + 1 2]+ x ln p +(n − x)ln(1− p) .)lnn − xnBy exp<strong>and</strong>ing the argument of the exponential in terms of y = x − np, where1 ≪ y ≪ np <strong>and</strong> keeping only the dominant terms, it can be shown thatf(x) ≈ √ 1[1√ exp − 1 (x − np) 2 ],2πn p(1 − p) 2 np(1 − p)which is of Gaussian <strong>for</strong>m with µ = np <strong>and</strong> σ = √ np(1 − p).Thus we see that the value of the Gaussian probability density function f(x) isa good approximation to the probability of obtaining x successes in n trials. Thisapproximation is actually very good even <strong>for</strong> relatively small n. For example, ifn =10<strong>and</strong>p =0.6 then the Gaussian approximation to the binomial distributionis (30.105) with µ =10× 0.6 =6<strong>and</strong>σ = √ 10 × 0.6(1 − 0.6) = 1.549. Theprobability functions f(x) <strong>for</strong> the binomial <strong>and</strong> associated Gaussian distributions<strong>for</strong> these parameters are given in table 30.4, <strong>and</strong> it can be seen that the Gaussianapproximation is a good one.Strictly speaking, however, since the Gaussian distribution is continuous <strong>and</strong>the binomial distribution is discrete, we should use the integral of f(x) <strong>for</strong>theGaussian distribution in the calculation of approximate binomial probabilities.More specifically, we should apply a continuity correction so that the discreteinteger x in the binomial distribution becomes the interval [x − 0.5, x+0.5] in1186


30.9 IMPORTANT CONTINUOUS DISTRIBUTIONSthe Gaussian distribution. Explicitly,Pr(X = x) ≈ 1 ∫ x+0.5[σ √ exp − 1 ( u − µ) 2]du.2π x−0.5 2 σThe Gaussian approximation is particularly useful <strong>for</strong> estimating the binomialprobability that X lies between the (integer) values x 1 <strong>and</strong> x 2 ,Pr(x 1 15.5) = Pr Z> √ =Pr(Z>−1.06)18=Pr(Z


PROBABILITYStirling’s approximation <strong>for</strong> large x givesx! ≈ √ ( x) x2πxeimplying thatln x! ≈ ln √ 2πx + x ln x − x,which, on substituting into (30.115), yieldsln f(x) ≈−λ + x ln λ − (x ln x − x) − ln √ 2πx.Since we expect the Poisson distribution to peak around x = λ, we substituteɛ = x − λ to obtain{ [ (ln f(x) ≈−λ +(λ + ɛ) ln λ − ln λ 1+ ɛ )]}+(λ + ɛ) − ln √ 2π(λ + ɛ).λUsing the expansion ln(1 + z) =z − z 2 /2+···, we find( ) ɛln f(x) ≈ ɛ − (λ + ɛ)λ − ɛ22λ 2 − ln √ ( ) ɛ2πλ −λ − ɛ22λ 2≈− ɛ22λ − ln √ 2πλ,when only the dominant terms are retained, after using the fact that ɛ is of theorder of the st<strong>and</strong>ard deviation of x, i.e.o<strong>for</strong>derλ 1/2 . On exponentiating thisresult we obtainf(x) ≈ √ 1 exp[−2πλ](x − λ)2,2λwhich is the Gaussian distribution with µ = λ <strong>and</strong> σ 2 = λ.The larger the value of λ, the better is the Gaussian approximation to thePoisson distribution; the approximation is reasonable even <strong>for</strong> λ = 5, but λ ≥ 10is safer. As in the case of the Gaussian approximation to the binomial distribution,a continuity correction is necessary since the Poisson distribution is discrete.◮E-mail messages are received by an author at an average rate of one per hour. Find theprobability that in a day the author receives 24 messages or more.We first define the r<strong>and</strong>om variableX = number of messages received in a day.Thus E[X] =1× 24 = 24, <strong>and</strong> so X ∼ Po(24). Since λ>10 we may approximate thePoisson distribution by X ∼ N(24, 24). Now the st<strong>and</strong>ard variable isZ = X √ − 24 ,24<strong>and</strong>, using the continuity correction, we find()23.5 − 24Pr(X >23.5) = Pr Z> √24=Pr(Z>−0.102) = Pr(Z


30.9 IMPORTANT CONTINUOUS DISTRIBUTIONSIn fact, almost all probability distributions tend towards a Gaussian when thenumbers involved become large – that this should happen is required by thecentral limit theorem, which we discuss in section 30.10.Multiple Gaussian distributionsSuppose X <strong>and</strong> Y are independent Gaussian-distributed r<strong>and</strong>om variables, sothat X ∼ N(µ 1 ,σ1 2)<strong>and</strong>Y ∼ N(µ 2,σ2 2 ). Let us now consider the r<strong>and</strong>om variableZ = X + Y . The PDF <strong>for</strong> this r<strong>and</strong>om variable may be found directly using(30.61), but it is easier to use the MGF. From (30.114), the MGFs of X <strong>and</strong> YareM X (t) =exp ( µ 1 t + 1 2 σ2 1t 2) , M Y (t) =exp ( µ 2 t + 1 2 σ2 2t 2) .Using (30.89), since X <strong>and</strong> Y are independent RVs, the MGF of Z = X + Y issimply the product of M X (t) <strong>and</strong>M Y (t). Thus, we haveM Z (t) =M X (t)M Y (t) =exp ( µ 1 t + 1 2 σ2 1t 2) exp ( µ 2 t + 1 2 σ2 2t 2)=exp [ (µ 1 + µ 2 )t + 1 2 (σ2 1 + σ2)t 2 2] ,which we recognise as the MGF <strong>for</strong> a Gaussian with mean µ 1 + µ 2 <strong>and</strong> varianceσ 2 1 + σ2 2 . Thus, Z is also Gaussian distributed: Z ∼ N(µ 1 + µ 2 , σ 2 1 + σ2 2 ).A similar calculation may be per<strong>for</strong>med to calculate the PDF of the r<strong>and</strong>omvariable W = X − Y . If we introduce the variable Ỹ = −Y then W = X + Ỹ ,where Ỹ ∼ N(−µ 1 , σ 2 1 ). Thus, using the result above, we find W ∼ N(µ 1 −µ 2 , σ 2 1 + σ2 2 ).◮An executive travels home from her office every evening. Her journey consists of a trainride, followed by a bicycle ride. The time spent on the train is Gaussian distributed withmean 52 minutes <strong>and</strong> st<strong>and</strong>ard deviation 1.8 minutes, while the time <strong>for</strong> the bicycle journeyis Gaussian distributed with mean 8 minutes <strong>and</strong> st<strong>and</strong>ard deviation 2.6 minutes. Assumingthese two factors are independent, estimate the percentage of occasions on which the wholejourney takes more than 65 minutes.We first define the r<strong>and</strong>om variablesX =timespentontrain, Y = time spent on bicycle,so that X ∼ N(52, (1.8) 2 )<strong>and</strong>Y ∼ N(8, (2.6) 2 ). Since X <strong>and</strong> Y are independent, the totaljourney time T = X + Y is distributed asT ∼ N(52 + 8, (1.8) 2 +(2.6) 2 )=N(60, (3.16) 2 ).The st<strong>and</strong>ard variable is thusZ = T − 603.16 ,<strong>and</strong> the required probability is given by()65 − 60Pr(T >65) = Pr Z> =Pr(Z>1.58) = 1 − 0.943 = 0.057.3.16Thus the total journey time exceeds 65 minutes on 5.7% of occasions. ◭1189


PROBABILITYThe above results may be extended. For example, if the r<strong>and</strong>om variablesX i , i =1, 2,...,n, are distributed as X i ∼ N(µ i ,σi 2 ) then the r<strong>and</strong>om variableZ = ∑ i c iX i (where the c i are constants) is distributed as Z ∼ N( ∑ i c iµ i , ∑ i c2 i σ2 i ).30.9.2 The log-normal distributionIf the r<strong>and</strong>om variable X follows a Gaussian distribution then the variableY = e X is described by a log-normal distribution. Clearly, if X can take valuesin the range −∞ to ∞, thenY will lie between 0 <strong>and</strong> ∞. The probability densityfunction <strong>for</strong> Y is found using the result (30.58). It isg(y) =f(x(y))dx∣ dy ∣ = 1 []σ √ 12π y exp (ln y − µ)2−2σ 2 .We note that µ <strong>and</strong> σ 2 are not the mean <strong>and</strong> variance of the log-normaldistribution, but rather the parameters of the corresponding Gaussian distribution<strong>for</strong> X. The mean <strong>and</strong> variance of Y , however, can be found straight<strong>for</strong>wardlyusing the MGF of X, which reads M X (t) =E[e tX ]=exp(µt + 1 2 σ2 t 2 ). Thus, themean of Y is given by<strong>and</strong> the variance of Y readsE[Y ]=E[e X ]=M X (1) = exp(µ + 1 2 σ2 ),V [Y ]=E[Y 2 ] − (E[Y ]) 2 = E[e 2X ] − (E[e X ]) 2= M X (2) − [M X (1)] 2 =exp(2µ + σ 2 )[exp(σ 2 ) − 1].In figure 30.15, we plot some examples of the log-normal distribution <strong>for</strong> variousvalues of the parameters µ <strong>and</strong> σ 2 .30.9.3 The exponential <strong>and</strong> gamma distributionsThe exponential distribution with positive parameter λ is given by{λe−λx<strong>for</strong> x>0,f(x) =(30.116)0 <strong>for</strong> x ≤ 0<strong>and</strong> satisfies ∫ ∞−∞f(x) dx = 1 as required. The exponential distribution occurs naturallyif we consider the distribution of the length of intervals between successiveevents in a Poisson process or, equivalently, the distribution of the interval (i.e.the waiting time) be<strong>for</strong>e the first event. If the average number of events per unitinterval is λ then on average there are λx events in interval x, so that from thePoisson distribution the probability that there will be no events in this interval isgiven byPr(no events in interval x) =e −λx .1190


30.9 IMPORTANT CONTINUOUS DISTRIBUTIONSg(y)10.80.6µ =0,σ =0µ =0,σ =0.5µ =0,σ =1.5µ =1,σ =10.40.20012 3 4yFigure 30.15 The PDF g(y) <strong>for</strong> the log-normal distribution <strong>for</strong> various valuesof the parameters µ <strong>and</strong> σ.The probability that an event occurs in the next infinitestimal interval [x, x + dx]is given by λdx,sothatPr(the first event occurs in interval [x, x + dx]) = e −λx λdx.Hence the required probability density function is given byf(x) =λe −λx .The expectation <strong>and</strong> variance of the exponential distribution can be evaluated as1/λ <strong>and</strong> (1/λ) 2 respectively. The MGF is given byM(t) =λλ − t . (30.117)We may generalise the above discussion to obtain the PDF <strong>for</strong> the intervalbetween every rth event in a Poisson process or, equivalently, the interval (waitingtime) be<strong>for</strong>e the rth event. We begin by using the Poisson distribution to give−λx (λx)r−1Pr(r − 1 events occur in interval x) =e(r − 1)! ,from which we obtain−λx (λx)r−1Pr(rth event occurs in the interval [x, x + dx]) = e(r − 1)! λdx.Thus the required PDF isλf(x) =(r − 1)! (λx)r−1 e −λx , (30.118)which is known as the gamma distribution of order r with parameter λ. Althoughour derivation applies only when r is a positive integer, the gamma distribution is1191


PROBABILITYf(x)10.80.6r =10.40.2r =2r =5r =1000 2 4 6 8 10 12 14 16 18 20xFigure 30.16 The PDF f(x) <strong>for</strong> the gamma distributions γ(λ, r) withλ =1<strong>and</strong> r =1, 2, 5, 10.defined <strong>for</strong> all positive r by replacing (r −1)! by Γ(r) in (30.118); see the appendix<strong>for</strong> a discussion of the gamma function Γ(x). If a r<strong>and</strong>om variable X is describedby a gamma distribution of order r with parameter λ, wewriteX ∼ γ(λ, r);we note that the exponential distribution is the special case γ(λ, 1). The gammadistribution γ(λ, r) is plotted in figure 30.16 <strong>for</strong> λ = 1 <strong>and</strong> r =1, 2, 5, 10. Forlarge r, the gamma distribution tends to the Gaussian distribution whose mean<strong>and</strong> variance are specified by (30.120) below.The MGF <strong>for</strong> the gamma distribution is obtained from that <strong>for</strong> the exponentialdistribution, by noting that we may consider the interval between every rth eventin a Poisson process as the sum of r intervals between successive events. Thus therth-order gamma variate is the sum of r independent exponentially distributedr<strong>and</strong>om variables. From (30.117) <strong>and</strong> (30.90), the MGF of the gamma distributionis there<strong>for</strong>e given by( ) r λM(t) = , (30.119)λ − tfrom which the mean <strong>and</strong> variance are found to beE[X] = r λ , V[X] = r λ 2 . (30.120)We may also use the above MGF to prove another useful theorem regardingmultiple gamma distributions. If X i ∼ γ(λ, r i ), i =1, 2,...,n, are independentgamma variates then the r<strong>and</strong>om variable Y = X 1 + X 2 + ···+ X n has MGFn∏( ) λri( ) λ r1+r 2+···+r nM(t) ==. (30.121)λ − t λ − ti=1Thus Y is also a gamma variate, distributed as Y ∼ γ(λ, r 1 + r 2 + ···+ r n ).1192


30.9 IMPORTANT CONTINUOUS DISTRIBUTIONS30.9.4 The chi-squared distributionIn subsection 30.6.2, we showed that if X is Gaussian distributed with mean µ <strong>and</strong>variance σ 2 , such that X ∼ N(µ, σ 2 ), then the r<strong>and</strong>om variable Y =(x − µ) 2 /σ 2is distributed as the gamma distribution Y ∼ γ( 1 2 , 1 2). Let us now consider nindependent Gaussian r<strong>and</strong>om variables X i ∼ N(µ i ,σi 2 ), i =1, 2,...,n, <strong>and</strong> definethe new variableχ 2 n =n∑ (X i − µ i ) 2. (30.122)i=1Using the result (30.121) <strong>for</strong> multiple gamma distributions, χ 2 n must be distributedas the gamma variate χ 2 n ∼ γ( 1 2 , 1 2n), which from (30.118) has the PDF1f(χ 2 2n)=Γ( 1 2 n)(1 2 χ2 n) (n/2)−1 exp(− 1 2 χ2 n)1=2 n/2 Γ( 1 n) (n/2)−1 exp(− 12 n)(χ2 2 χ2 n). (30.123)This is known as the chi-squared distribution of order n <strong>and</strong> has numerousapplications in statistics (see chapter 31). Setting λ = 1 2 <strong>and</strong> r = 1 2n in (30.120),we find thatE[χ 2 n]=n,σ 2 iV [χ 2 n]=2n.An important generalisation occurs when the n Gaussian variables X i are notlinearly independent but are instead required to satisfy a linear constraint of the<strong>for</strong>mc 1 X 1 + c 2 X 2 + ···+ c n X n =0, (30.124)in which the constants c i are not all zero. In this case, it may be shown (seeexercise 30.40) that the variable χ 2 n defined in (30.122) is still described by a chisquareddistribution, but one of order n − 1. Indeed, this result may be triviallyextended to show that if the n Gaussian variables X i satisfy m linear constraintsof the <strong>for</strong>m (30.124) then the variable χ 2 n defined in (30.122) is described by achi-squared distribution of order n − m.30.9.5 The Cauchy <strong>and</strong> Breit–Wigner distributionsA r<strong>and</strong>om variable X (in the range −∞ to ∞) that obeys the Cauchy distributionis described by the PDFf(x) = 1 π119311+x 2 .


PROBABILITYf(x)0.80.6x 0 =0,x 0 =2,Γ=1 Γ=10.40.20x 0 =0,Γ=3−4 −2 0 2 4xFigure 30.17 The PDF f(x) <strong>for</strong> the Breit–Wigner distribution <strong>for</strong> differentvalues of the parameters x 0 <strong>and</strong> Γ.This is a special case of the Breit–Wigner distributionf(x) = 1 π12 Γ14 Γ2 +(x − x 0 ) , 2which is encountered in the study of nuclear <strong>and</strong> particle physics. In figure 30.17,we plot some examples of the Breit–Wigner distribution <strong>for</strong> several values of theparameters x 0 <strong>and</strong> Γ.We see from the figure that the peak (or mode) of the distribution occursat x = x 0 . It is also straight<strong>for</strong>ward to show that the parameter Γ is equal tothe width of the peak at half the maximum height. Although the Breit–Wignerdistribution is symmetric about its peak, it does not <strong>for</strong>mally possess a mean sincethe integrals ∫ 0−∞ xf(x) dx <strong>and</strong> ∫ ∞0xf(x) dx both diverge. Similar divergences occur<strong>for</strong> all higher moments of the distribution.30.9.6 The uni<strong>for</strong>m distributionFinally we mention the very simple, but common, uni<strong>for</strong>m distribution, whichdescribes a continuous r<strong>and</strong>om variable that has a constant PDF over its allowedrange of values. If the limits on X are a <strong>and</strong> b then{1/(b − a) <strong>for</strong> a ≤ x ≤ b,f(x) =0 otherwise.The MGF of the uni<strong>for</strong>m distribution is found to beM(t) = ebt − e at(b − a)t ,1194


30.10 THE CENTRAL LIMIT THEOREM<strong>and</strong> its mean <strong>and</strong> variance are given byE[X] = a + b2− a)2, V[X] =(b .1230.10 The central limit theoremIn subsection 30.9.1 we discussed approximating the binomial <strong>and</strong> Poisson distributionsby the Gaussian distribution when the number of trials is large. We nowdiscuss why the Gaussian distribution is so common <strong>and</strong> there<strong>for</strong>e so important.The central limit theorem may be stated as follows.Central limit theorem. Suppose that X i , i =1, 2,...,n,areindependent r<strong>and</strong>omvariables, each of which is described by a probability density function f i (x) (thesemay all be different) with a mean µ i <strong>and</strong> a variance σi 2 . The r<strong>and</strong>om variable Z =(∑i X i)/n, i.e. the ‘mean’ of the Xi , has the following properties:(i) its expectation value is given by E[Z] = (∑ i µ i)/n;(ii) its variance is given by V [Z] = (∑ )i σ2 i /n 2 ;(iii) as n →∞the probability function of Z tends to a Gaussian with correspondingmean <strong>and</strong> variance.We note that <strong>for</strong> the theorem to hold, the probability density functions f i (x)must possess <strong>for</strong>mal means <strong>and</strong> variances. Thus, <strong>for</strong> example, if any of the X iwere described by a Cauchy distribution then the theorem would not apply.Properties (i) <strong>and</strong> (ii) of the theorem are easily proved, as follows. FirstlyE[Z] = 1 n (E[X 1]+E[X 2 ]+···+ E[X n ]) = 1 ∑n (µ i1 + µ 2 + ···+ µ n )=µ in ,a result which does not require that the X i are independent r<strong>and</strong>om variables. Ifµ i = µ <strong>for</strong> all i then this becomesE[Z] = nµ n = µ.Secondly, if the X i are independent, it follows from an obvious extension of(30.68) that[ 1V [Z] =Vn (X 1 + X 2 + ···+ X n )]= 1 ∑n (V 2 [X i1]+V [X 2 ]+···+ V [X n ]) =σ2 in 2 .Let us now consider property (iii), which is the reason <strong>for</strong> the ubiquity ofthe Gaussian distribution <strong>and</strong> is most easily proved by considering the momentgenerating function M Z (t) ofZ. From (30.90), this MGF is given byn∏( ) tM Z (t) = M Xi ,ni=11195


PROBABILITYwhere M Xi (t) istheMGFoff i (x). Now( ) tM Xi =1+ t n n E[X i]+ 1 t 22n 2 E[X2 i ]+···t=1+µ in + 1 2 (σ2 i + µ 2 i ) t2n 2 + ··· ,<strong>and</strong> as n becomes large( ) ( t µi tM Xi ≈ expnn + 1 t 2 )2 σ2 in 2 ,as may be verified by exp<strong>and</strong>ing the exponential up to terms including (t/n) 2 .There<strong>for</strong>en∏(µi tM Z (t) ≈ expn + 1 t 2 ) (∑2 σ2 iin 2 =expµ ∑ )in t + 1 i σ2 i2n 2 t 2 .i=1Comparing this with the <strong>for</strong>m of the MGF <strong>for</strong> a Gaussian distribution, (30.114),we can see that the probability density function g(z) ofZ tends to a Gaussian distributionwith mean ∑ i µ i/n <strong>and</strong> variance ∑ i σ2 i /n2 . In particular, if we considerZ to be the mean of n independent measurements of the same r<strong>and</strong>om variable X(so that X i = X <strong>for</strong> i =1, 2,...,n) then, as n →∞, Z has a Gaussian distributionwith mean µ <strong>and</strong> variance σ 2 /n.We may use the central limit theorem to derive an analogous result to (iii)above <strong>for</strong> the product W = X 1 X 2 ···X n of the n independent r<strong>and</strong>om variablesX i . Provided the X i only take values between zero <strong>and</strong> infinity, we may writeln W =lnX 1 +lnX 2 + ···+lnX n ,which is simply the sum of n new r<strong>and</strong>om variables ln X i . Thus, provided thesenew variables each possess a <strong>for</strong>mal mean <strong>and</strong> variance, the PDF of ln W willtend to a Gaussian in the limit n →∞, <strong>and</strong> so the product W will be describedby a log-normal distribution (see subsection 30.9.2).30.11 Joint distributionsAs mentioned briefly in subsection 30.4.3, it is common in the physical sciences toconsider simultaneously two or more r<strong>and</strong>om variables that are not independent,in general, <strong>and</strong> are thus described by joint probability density functions. We willreturn to the subject of the interdependence of r<strong>and</strong>om variables after firstpresenting some of the general ways of characterising joint distributions. Wewill concentrate mainly on bivariate distributions, i.e. distributions of only twor<strong>and</strong>om variables, though the results may be extended readily to multivariatedistributions. The subject of multivariate distributions is large <strong>and</strong> a detailedstudy is beyond the scope of this book; the interested reader should there<strong>for</strong>e1196


30.11 JOINT DISTRIBUTIONSconsult one of the many specialised texts. However, we do discuss the multinomial<strong>and</strong> multivariate Gaussian distributions, in section 30.15.The first thing to note when dealing with bivariate distributions is that thedistinction between discrete <strong>and</strong> continuous distributions may not be as clear as<strong>for</strong> the single variable case; the r<strong>and</strong>om variables can both be discrete, or bothcontinuous, or one discrete <strong>and</strong> the other continuous. In general, <strong>for</strong> the r<strong>and</strong>omvariables X <strong>and</strong> Y , the joint distribution will take an infinite number of valuesunless both X <strong>and</strong> Y have only a finite number of values. In this chapter wewill consider only the cases where X <strong>and</strong> Y are either both discrete or bothcontinuous r<strong>and</strong>om variables.30.11.1 Discrete bivariate distributionsIn direct analogy with the one-variable (univariate) case, if X is a discrete r<strong>and</strong>omvariable that takes the values {x i } <strong>and</strong> Y one that takes the values {y j } then theprobability function of the joint distribution is defined as{Pr(X = xi , Y = y j ) <strong>for</strong> x = x i , y = y j ,f(x, y) =0 otherwise.We may there<strong>for</strong>e think of f(x, y) as a set of spikes at valid points in the xy-plane,whose height at (x i ,y i ) represents the probability of obtaining X = x i <strong>and</strong> Y = y j .The normalisation of f(x, y) implies∑ ∑f(x i ,y j )=1, (30.125)ijwhere the sums over i <strong>and</strong> j take all valid pairs of values. We can also define thecumulative probability functionF(x, y) = ∑ ∑f(x i ,y j ), (30.126)x i≤x y j≤yfrom which it follows that the probability that X lies in the range [a 1 ,a 2 ]<strong>and</strong>Ylies in the range [b 1 ,b 2 ] is given byPr(a 1


PROBABILITY30.11.2 Continuous bivariate distributionsIn the case where both X <strong>and</strong> Y are continuous r<strong>and</strong>om variables, the PDF ofthe joint distribution is defined byf(x, y) dx dy =Pr(x


30.12 PROPERTIES OF JOINT DISTRIBUTIONS30.11.3 Marginal <strong>and</strong> conditional distributionsGiven a bivariate distribution f(x, y), we may be interested only in the probabilityfunction <strong>for</strong> X irrespective of the value of Y (or vice versa). This marginaldistribution of X is obtained by summing or integrating, as appropriate, thejoint probability distribution over all allowed values of Y . Thus, the marginaldistribution of X (<strong>for</strong> example) is given by{∑jf X (x) =f(x, y j) <strong>for</strong> a discrete distribution,∫ (30.130)f(x, y) dy <strong>for</strong> a continuous distribution.It is clear that an analogous definition exists <strong>for</strong> the marginal distribution of Y .Alternatively, one might be interested in the probability function of X giventhat Y takes some specific value of Y = y 0 ,i.e.Pr(X = x|Y = y 0 ). This conditionaldistribution of X is given byg(x) = f(x, y 0)f Y (y 0 ) ,where f Y (y) is the marginal distribution of Y . The division by f Y (y 0 ) is necessaryin order that g(x) is properly normalised.30.12 Properties of joint distributionsThe probability density function f(x, y) contains all the in<strong>for</strong>mation on the jointprobability distribution of two r<strong>and</strong>om variables X <strong>and</strong> Y . In a similar mannerto that presented <strong>for</strong> univariate distributions, however, it is conventional tocharacterise f(x, y) by certain of its properties, which we now discuss. Onceagain, most of these properties are based on the concept of expectation values,which are defined <strong>for</strong> joint distributions in an analogous way to those <strong>for</strong> singlevariabledistributions (30.46). Thus, the expectation value of any function g(X,Y )of the r<strong>and</strong>om variables X <strong>and</strong> Y is given by{∑ ∑i jE[g(X,Y )] =g(x i,y j )f(x i ,y j ) <strong>for</strong> the discrete case,g(x, y)f(x, y) dx dy <strong>for</strong> the continuous case.∫ ∞ ∫ ∞−∞ −∞30.12.1 MeansThe means of X <strong>and</strong> Y are defined respectively as the expectation values of thevariables X <strong>and</strong> Y . Thus, the mean of X is given by{∑ ∑i jE[X] =µ X =x if(x i ,y j ) <strong>for</strong> the discrete case,∫ ∞ ∫ ∞−∞ −∞ xf(x, y) dx dy <strong>for</strong> the continuous case. (30.131)E[Y ] is obtained in a similar manner.1199


PROBABILITY◮Show that if X <strong>and</strong> Y are independent r<strong>and</strong>om variables then E[XY ]=E[X]E[Y ].LetusconsiderthecasewhereX <strong>and</strong> Y are continuous r<strong>and</strong>om variables. Since X <strong>and</strong>Y are independent f(x, y) =f X (x)f Y (y), so thatE[XY ]=∫ ∞ ∫ ∞xyf X (x)f Y (y) dx dy =∫ ∞xf X (x) dx∫ ∞−∞ −∞−∞−∞An analogous proof exists <strong>for</strong> the discrete case. ◭yf Y (y) dy = E[X]E[Y ].30.12.2 VariancesThe definitions of the variances of X <strong>and</strong> Y are analogous to those <strong>for</strong> thesingle-variable case (30.48), i.e. the variance of X is given by{∑ ∑V [X] =σX 2 = i j (x i − µ X ) 2 f(x i ,y j ) <strong>for</strong> the discrete case,∫ ∞ ∫ ∞−∞ −∞ (x − µ X) 2 f(x, y) dx dy <strong>for</strong> the continuous case. (30.132)Equivalent definitions exist <strong>for</strong> the variance of Y .30.12.3 Covariance <strong>and</strong> correlationMeans <strong>and</strong> variances of joint distributions provide useful in<strong>for</strong>mation abouttheir marginal distributions, but we have not yet given any indication of how tomeasure the relationship between the two r<strong>and</strong>om variables. Of course, it maybe that the two r<strong>and</strong>om variables are independent, but often this is not so. Forexample, if we measure the heights <strong>and</strong> weights of a sample of people we wouldnot be surprised to find a tendency <strong>for</strong> tall people to be heavier than short people<strong>and</strong> vice versa. We will show in this section that two functions, the covariance<strong>and</strong> the correlation, can be defined <strong>for</strong> a bivariate distribution <strong>and</strong> that these areuseful in characterising the relationship between the two r<strong>and</strong>om variables.The covariance of two r<strong>and</strong>om variables X <strong>and</strong> Y is defined byCov[X,Y ]=E[(X − µ X )(Y − µ Y )], (30.133)where µ X <strong>and</strong> µ Y are the expectation values of X <strong>and</strong> Y respectively. Clearlyrelated to the covariance is the correlation of the two r<strong>and</strong>om variables, definedbyCorr[X,Y ]= Cov[X,Y ] , (30.134)σ X σ Ywhere σ X <strong>and</strong> σ Y are the st<strong>and</strong>ard deviations of X <strong>and</strong> Y respectively. It can beshown that the correlation function lies between −1 <strong>and</strong> +1. If the value assumedis negative, X <strong>and</strong> Y are said to be negatively correlated, if it is positive they aresaid to be positively correlated <strong>and</strong>ifitiszerotheyaresaidtobeuncorrelated.We will now justify the use of these terms.1200


30.12 PROPERTIES OF JOINT DISTRIBUTIONSOne particularly useful consequence of its definition is that the covarianceof two independent variables, X <strong>and</strong> Y , is zero. It immediately follows from(30.134) that their correlation is also zero, <strong>and</strong> this justifies the use of the term‘uncorrelated’ <strong>for</strong> two such variables. To show this extremely important propertywe first note thatCov[X,Y ]=E[(X − µ X )(Y − µ Y )]= E[XY − µ X Y − µ Y X + µ X µ Y ]= E[XY ] − µ X E[Y ] − µ Y E[X]+µ X µ Y= E[XY ] − µ X µ Y . (30.135)Now, if X <strong>and</strong> Y are independent then E[XY ]=E[X]E[Y ]=µ X µ Y <strong>and</strong> soCov[X,Y ] = 0. It is important to note that the converse of this result is notnecessarily true; two variables dependent on each other can still be uncorrelated.In other words, it is possible (<strong>and</strong> not uncommon) <strong>for</strong> two variables X <strong>and</strong> Yto be described by a joint distribution f(x, y) thatcannot be factorised into aproduct of the <strong>for</strong>m g(x)h(y), but <strong>for</strong> which Corr[X,Y ] = 0. Indeed, from thedefinition (30.133), we see that <strong>for</strong> any joint distribution f(x, y) that is symmetricin x about µ X (or similarly in y) we have Corr[X,Y ]=0.We have already asserted that if the correlation of two r<strong>and</strong>om variables ispositive (negative) they are said to be positively (negatively) correlated. We havealso stated that the correlation lies between −1 <strong>and</strong> +1. The terminology suggeststhat if the two RVs are identical (i.e. X = Y ) then they are completely correlated<strong>and</strong> that their correlation should be +1. Likewise, if X = −Y then the functionsare completely anticorrelated <strong>and</strong> their correlation should be −1. Values of thecorrelation function between these extremes show the existence of some degreeof correlation. In fact it is not necessary that X = Y <strong>for</strong> Corr[X,Y ] = 1; it issufficient that Y is a linear function of X, i.e.Y = aX + b (with a positive). If ais negative then Corr[X,Y ]=−1. To show this we first note that µ Y = aµ X + b.NowY = aX + b = aX + µ Y − aµ X ⇒ Y − µ Y = a(X − µ X ),<strong>and</strong> so using the definition of the covariance (30.133)Cov[X,Y ]=aE[(X − µ X ) 2 ]=aσX.2It follows from the properties of the variance (subsection 30.5.3) that σ Y = |a|σ X<strong>and</strong> so, using the definition (30.134) of the correlation,Corr[X,Y ]= aσ2 X|a|σX2 = a|a| ,which is the stated result.It should be noted that, even if the possibilities of X <strong>and</strong> Y being non-zero aremutually exclusive, Corr[X,Y ] need not have value ±1.1201


PROBABILITY◮A biased die gives probabilities 1 p, p, p, p, p, 2p of throwing 1, 2, 3, 4, 5, 6 respectively.2If the r<strong>and</strong>om variable X is the number shown on the die <strong>and</strong> the r<strong>and</strong>om variable Y isdefined as X 2 , calculate the covariance <strong>and</strong> correlation of X <strong>and</strong> Y .We have already calculated in subsections 30.2.1 <strong>and</strong> 30.5.4 thatp = 253, E[X] =13Using (30.135), we obtainNow E[X 3 ]isgivenby13 , E[ X 2] = 25313, V[X] =480169 .Cov[X,Y ]=Cov[X,X 2 ]=E[X 3 ] − E[X]E[X 2 ].E[X 3 ]=1 3 × 1 p 2 +(23 +3 3 +4 3 +5 3 )p +6 3 × 2p= 13132 p = 101,<strong>and</strong> the covariance of X <strong>and</strong> Y is given byCov[X,Y ] = 101 − 5313 × 25313 = 3660169 .The correlation is defined by Corr[X,Y ]=Cov[X,Y ]/σ X σ Y . The st<strong>and</strong>ard deviation ofY may be calculated from the definition of the variance. Letting µ Y = E[X 2 ]= 253 gives13σY 2 = p ( )1 2 2 ( )− µ Y + p 2 2 2 ( )− µ Y + p 3 2 2 ( )− µ Y + p 4 2 2− µ Y2+ p ( )5 2 2 ( )− µ Y +2p 6 2 2− µ Y187 356 28 824= p =169 169 .We deduce thatCorr[X,Y ]= 3660√ √169 169169 28 824 480 ≈ 0.984.Thus the r<strong>and</strong>om variables X <strong>and</strong> Y display a strong degree of positive correlation, as wewould expect. ◭We note that the covariance of X <strong>and</strong> Y occurs in various expressions. Forexample, if X <strong>and</strong> Y are not independent thenV [X + Y ]=E [ (X + Y ) 2] − (E[X + Y ]) 2= E [ X 2] +2E[XY ]+E [ Y 2] −{(E[X]) 2 +2E[X]E[Y ]+(E[Y ]) 2 }= V [X]+V [Y ]+2(E[XY ] − E[X]E[Y ])= V [X]+V [Y ]+2Cov[X,Y ].1202


30.12 PROPERTIES OF JOINT DISTRIBUTIONSMore generally, we find (<strong>for</strong> a, b <strong>and</strong> c constant)V [aX + bY + c] =a 2 V [X]+b 2 V [Y ]+2ab Cov[X,Y ].(30.136)Note that if X <strong>and</strong> Y are in fact independent then Cov[X,Y ] = 0 <strong>and</strong> we recoverthe expression (30.68) in subsection 30.6.4.We may use (30.136) to obtain an approximate expression <strong>for</strong> V [ f(X,Y )]<strong>for</strong> any arbitrary function f, even when the r<strong>and</strong>om variables X <strong>and</strong> Y arecorrelated. Approximating f(X,Y ) by the linear terms of its Taylor expansionabout the point (µ X ,µ Y ), we have( )( )∂f∂ff(X,Y ) ≈ f(µ X ,µ Y )+ (X − µ X )+ (Y − µ Y ),∂X∂Y(30.137)where the partial derivatives are evaluated at X = µ X <strong>and</strong> Y = µ Y . Taking thevariance of both sides, <strong>and</strong> using (30.136), we find( ) 2 ( ) 2 ( )( )∂f∂f∂f ∂fV [ f(X,Y )] ≈ V [X]+ V [Y ]+2Cov[X,Y ].∂X∂Y∂X ∂Y(30.138)Clearly, if Cov[X,Y ] = 0, we recover the result (30.69) derived in subsection 30.6.4.We note that (30.138) is exact if f(X,Y ) is linear in X <strong>and</strong> Y .For several variables X i , i =1, 2,...,n, we can define the symmetric (positivedefinite) covariance matrix whose elements are<strong>and</strong> the symmetric (positive definite) correlation matrixV ij =Cov[X i ,X j ], (30.139)ρ ij =Corr[X i ,X j ].The diagonal elements of the covariance matrix are the variances of the variables,whilst those of the correlation matrix are unity. For several variables, (30.138)generalises toV [f(X 1 ,X 2 ,...,X n )] ≈ ∑ i( ∂f∂X i) 2V [X i ]+ ∑ i∑j≠i( ∂f∂X i)( ∂f∂X j)Cov[X i ,X j ],where the partial derivatives are evaluated at X i = µ Xi .1203


PROBABILITY◮A card is drawn at r<strong>and</strong>om from a normal 52-card pack <strong>and</strong> its identity noted. The cardis replaced, the pack shuffled <strong>and</strong> the process repeated. R<strong>and</strong>om variables W, X, Y, Z aredefined as follows:W =2 if the drawn card is a heart; W =0otherwise.X =4 if the drawn card is an ace, king, or queen; X =2if the card isajackorten;X =0otherwise.Y =1 if the drawn card is red; Y =0otherwise.Z =2 if the drawn card is black <strong>and</strong> an ace, king or queen; Z =0otherwise.Establish the correlation matrix <strong>for</strong> W, X, Y, Z.The means of the variables are given byµ W =2× 1 = 1 , µ 4 2 X = ( ) ( )4 × 313 + 2 ×213 =16, 13µ Y =1× 1 = 1 , µ 2 2 Z =2× 6 = 3 . 52 13The variances, calculated from σU 2 = V [U] =E [ U 2] − (E[U]) 2 ,whereU = W , X, Y orZ, areσW 2 = ( ) (4 × 1 4 − 1) 22 =3, 4 σ2 X = ( 16 × 313σ 2 Y = ( 1 × 1 2)−( 12) 2=14 , σ2 Z = ( 4 × 652) ( ) (+ 4 ×2 − 16)−( 31313) 2=69169 .13) 2=472169 ,The covariances are found by first calculating E[WX] etc. <strong>and</strong> then <strong>for</strong>ming E[WX]−µ W µ Xetc.E[WX]=2(4) ( ) (352 +2(2) 2)52 =8, Cov[W,X]= 8 − ( 1 16)13 13 2 13 =0,E[WY] = 2(1) ( )14 =1, Cov[W,Y]= 1 − ( 1 1)2 2 2 2 =1, 4(E[WZ]=0, Cov[W,Z]=0− 1 3)2 13 = −3, 26E[XY ] = 4(1) ( ) (652 +2(1) 4)52 =88, Cov[X,Y ]= − ( 16 1)13 13 13 2 =0,E[XZ] = 4(2) ( 652)=1212, Cov[X,Z]= − 1613 13 13( 3)13 =108, 169(E[YZ]=0, Cov[Y,Z]=0− 1 3)2 13 = −3. 26The correlations Corr[W,X] <strong>and</strong> Corr[X,Y ] are clearly zero; the remainder are given by(Corr[W,Y]= 1 3 × ) 1 −1/24 4 4 = 0.577,(Corr[W,Z]=− 3 3 × ) 69 −1/226 4 169 = −0.209,Corr[X,Z]= 108169Corr[Y,Z]=− 326( 472169 × 69) −1/2169 = 0.598,( 1 × ) 69 −1/24 169 = −0.361.Finally, then, we can write down the correlation matrix:⎛⎞1 0 0.58 −0.21⎜ 0 1 0 0.60 ⎟ρ = ⎝0.58 0 1 −0.36⎠ .−0.21 0.60 −0.36 11204


30.13 GENERATING FUNCTIONS FOR JOINT DISTRIBUTIONSAs would be expected, X is uncorrelated with either W or Y , colour <strong>and</strong> face-value beingtwo independent characteristics. Positive correlations are to be expected between W <strong>and</strong>Y <strong>and</strong> between X <strong>and</strong> Z; both correlations are fairly strong. Moderate anticorrelationsexist between Z <strong>and</strong> both W <strong>and</strong> Y , reflecting the fact that it is impossible <strong>for</strong> W <strong>and</strong> Yto be positive if Z is positive. ◭Finally, let us suppose that the r<strong>and</strong>om variables X i , i =1, 2,...,n, are relatedto a second set of r<strong>and</strong>om variables Y k = Y k (X 1 ,X 2 ,...,X n ), k =1, 2,...,m.Byexp<strong>and</strong>ing each Y k as a Taylor series as in (30.137) <strong>and</strong> inserting the resultingexpressions into the definition of the covariance (30.133), we find that the elementsof the covariance matrix <strong>for</strong> the Y k variables are given byCov[Y k ,Y l ] ≈ ∑ ∑( )( )∂Yk ∂YlCov[X i ,X j ].∂Xi j i ∂X j(30.140)It is straight<strong>for</strong>ward to show that this relation is exact if the Y k are linearcombinations of the X i . Equation (30.140) can then be written in matrix <strong>for</strong>m asV Y = SV X S T , (30.141)where V Y <strong>and</strong> V X are the covariance matrices of the Y k <strong>and</strong> X i variables respectively<strong>and</strong> S is the rectangular m × n matrix with elements S ki = ∂Y k /∂X i .30.13 Generating functions <strong>for</strong> joint distributionsIt is straight<strong>for</strong>ward to generalise the discussion of generating function in section30.7 to joint distributions. For a multivariate distribution f(X 1 ,X 2 ,...,X n ) ofnon-negative integer r<strong>and</strong>om variables X i , i =1, 2,...,n, we define the probabilitygenerating function to beΦ(t 1 ,t 2 ,...,t n )=E[t X11 tX2 2 ···tXn n ].As in the single-variable case, we may also define the closely related momentgenerating function, which has wider applicability since it is not restricted tonon-negative integer r<strong>and</strong>om variables but can be used with any set of discreteor continuous r<strong>and</strong>om variables X i (i =1, 2,...,n). The MGF of the multivariatedistribution f(X 1 ,X 2 ,...,X n ) is defined asM(t 1 ,t 2 ,...,t n )=E[e t1X1 e t2X2 ···e tnXn ]=E[e t1X1+t2X2+···+tnXn ](30.142)<strong>and</strong> may be used to evaluate (joint) moments of f(X 1 ,X 2 ,...,X n ). By per<strong>for</strong>minga derivation analogous to that presented <strong>for</strong> the single-variable case in subsection30.7.2, it can be shown that∂m1+m2+···+mn M(0, 0,...,0)E[X m11 Xm2 2···Xn mn]= . (30.143)∂t m11 ∂tm2 2 ···∂tmn n1205


PROBABILITYFinally we note that, by analogy with the single-variable case, the characteristicfunction <strong>and</strong> the cumulant generating function of a multivariate distribution aredefined respectively asC(t 1 ,t 2 ,...,t n )=M(it 1 ,it 2 ,...,it n ) <strong>and</strong> K(t 1 ,t 2 ,...,t n )=lnM(t 1 ,t 2 ,...,t n ).◮Suppose that the r<strong>and</strong>om variables X i , i =1, 2,...,n,aredescribedbythePDFf(x) =f(x 1 ,x 2 ,...,x n )=N exp(− 1 2 xT Ax),where the column vector x =(x 1 x 2 ··· x n ) T , A is an n × n symmetric matrix <strong>and</strong> Nis a normalisation constant such that∫∫ ∞ ∫ ∞ ∫ ∞f(x) d n x ≡ ··· f(x 1 ,x 2 ,...,x n ) dx 1 dx 2 ···dx n =1.∞Find the MGF of f(x).−∞−∞−∞From (30.142), the MGF is given by∫M(t 1 ,t 2 ,...,t n )=N exp(− 1 2 xT Ax + t T x) d n x, (30.144)∞where the column vector t =(t 1 t 2 ··· t n ) T . In order to evaluate this multiple integral,we begin by noting thatx T Ax − 2t T x =(x − A −1 t) T A(x − A −1 t) − t T A −1 t,which is the matrix equivalent of ‘completing the square’. Using this expression in (30.144)<strong>and</strong> making the substitution y = x − A −1 t,weobtainM(t 1 ,t 2 ,...,t n )=c exp( 1 2 tT A −1 t), (30.145)where the constant c is given by∫c = N exp(− 1 2 yT Ay) d n y.∞From the normalisation condition <strong>for</strong> N, weseethatc = 1, as indeed it must be in orderthat M(0, 0,...,0) = 1. ◭30.14 Trans<strong>for</strong>mation of variables in joint distributionsSuppose the r<strong>and</strong>om variables X i , i =1, 2,...,n, are described by the multivariatePDF f(x 1 ,x 2 ...,x n ). If we wish to consider r<strong>and</strong>om variables Y j , j =1, 2,...,m,related to the X i by Y j = Y j (X 1 ,X 2 ,...,X m ) then we may calculate g(y 1 ,y 2 ,...,y m ),the PDF <strong>for</strong> the Y j , in a similar way to that in the univariate case by dem<strong>and</strong>ingthat|f(x 1 ,x 2 ...,x n ) dx 1 dx 2 ···dx n | = |g(y 1 ,y 2 ,...,y m ) dy 1 dy 2 ···dy m |.From the discussion of changing the variables in multiple integrals given inchapter 6 it follows that, in the special case where n = m,g(y 1 ,y 2 ,...,y m )=f(x 1 ,x 2 ...,x n )|J|,1206


30.15 IMPORTANT JOINT DISTRIBUTIONSwhere∣ ∣∣∣∣∣∣∣∣∣ ∂x 1 ∂x∣n ∣∣∣∣∣∣∣∣∣...J ≡ ∂(x ∂y 1 ∂y 11,x 2 ...,x n )∂(y 1 ,y 2 ,...,y n ) = .. .. . ,∂x 1 ∂x n...∂y n ∂y nis the Jacobian of the x i with respect to the y j .◮Suppose that the r<strong>and</strong>om variables X i , i =1, 2,...,n, are independent <strong>and</strong> Gaussian distributedwith means µ i <strong>and</strong> variances σi2 respectively. Find the PDF <strong>for</strong> the new variablesZ i =(X i − µ i )/σ i , i =1, 2,...,n. By considering an elemental spherical shell in Z-space,find the PDF of the chi-squared r<strong>and</strong>om variable χ 2 n = ∑ ni=1 Z i 2 .Since the X i are independent r<strong>and</strong>om variables,[]1n∑ (x i − µ i ) 2f(x 1 ,x 2 ,...,x n )=f(x 1 )f(x 2 ) ···f(x n )=exp −.(2π) n/2 σ 1 σ 2 ···σ n 2σ 2 i=1 iTo derive the PDF <strong>for</strong> the variables Z i ,werequire|f(x 1 ,x 2 ,...,x n ) dx 1 dx 2 ···dx n | = |g(z 1 ,z 2 ,...,z n ) dz 1 dz 2 ···dz n |,<strong>and</strong>, noting that dz i = dx i /σ i ,weobtain)1g(z 1 ,z 2 ,...,z n )=(−(2π) exp 1 n∑z n/2 i2 .2i=1Let us now consider the r<strong>and</strong>om variable χ 2 n = ∑ ni=1 Z i 2 , which we may regard as thesquare of the distance from the origin in the n-dimensional Z-space. We now require thatg(z 1 ,z 2 ,...,z n ) dz 1 dz 2 ···dz n = h(χ 2 n)dχ 2 n.If we consider the infinitesimal volume dV = dz 1 dz 2 ···dz n to be that enclosed by then-dimensional spherical shell of radius χ n <strong>and</strong> thickness dχ n then we may write dV =dχ n , <strong>for</strong> some constant A. We thus obtainAχ n−1nh(χ 2 n)dχ 2 n ∝ exp(− 1 2 χ2 n)χ n−1n dχ n ∝ exp(− 1 2 χ2 n)χn n−2 dχ 2 n,where we have used the fact that dχ 2 n =2χ n dχ n . Thus we see that the PDF <strong>for</strong> χ 2 n is givenbyh(χ 2 n)=B exp(− 1 2 χ2 n)χ n−2n ,<strong>for</strong> some constant B. This constant may be determined from the normalisation condition∫ ∞h(χ 2 n) dχ 2 n =10<strong>and</strong> is found to be B = [2 n/2 Γ( 1 2 n)]−1 . This is the nth-order chi-squared distributiondiscussed in subsection 30.9.4. ◭30.15 Important joint distributionsIn this section we will examine two important multivariate distributions, themultinomial distribution, which is an extension of the binomial distribution, <strong>and</strong>the multivariate Gaussian distribution.1207


PROBABILITY30.15.1 The multinomial distributionThe binomial distribution describes the probability of obtaining x ‘successes’ fromn independent trials, where each trial has only two possible outcomes. This maybe generalised to the case where each trial has k possible outcomes with respectiveprobabilities p 1 , p 2 , ..., p k . If we consider the r<strong>and</strong>om variables X i , i =1, 2,...,n,to be the number of outcomes of type i in n trials then we may calculate theirjoint probability functionf(x 1 ,x 2 ,...,x k )=Pr(X 1 = x 1 ,X 2 = x 2 , ..., X k = x k ),wherewemusthave ∑ ki=1 x i = n. Inn trials the probability of obtaining x 1outcomes of type 1, followed by x 2 outcomes of type 2 etc. is given byp x11 px2 2 ···pxk k .However, the number of distinguishable permutations of this result isn!x 1 !x 2 ! ···x k ! ,<strong>and</strong> thusn!f(x 1 ,x 2 ,...,x k )=x 1 !x 2 ! ···x k ! px1 1 px2 2 ···pxk k . (30.146)This is the multinomial probability distribution.If k = 2 then the multinomial distribution reduces to the familiar binomialdistribution. Although in this <strong>for</strong>m the binomial distribution appears to be afunction of two r<strong>and</strong>om variables, it must be remembered that, in fact, sincep 2 =1− p 1 <strong>and</strong> x 2 = n − x 1 , the distribution of X 1 is entirely determined by theparameters p <strong>and</strong> n. ThatX 1 has a binomial distribution is shown by rememberingthat it represents the number of objects of a particular type obtained fromsampling with replacement, which led to the original definition of the binomialdistribution. In fact, any of the r<strong>and</strong>om variables X i has a binomial distribution,i.e. the marginal distribution of each X i is binomial with parameters n <strong>and</strong> p i .Itimmediately follows thatE[X i ]=np i <strong>and</strong> V [X i ] 2 = np i (1 − p i ). (30.147)◮At a village fête patrons were invited, <strong>for</strong> a 10 p entry fee, to pick without looking sixtickets from a drum containing equal large numbers of red, blue <strong>and</strong> green tickets. If fiveor more of the tickets were of the same colour a prize of 100 p was awarded. A consolationaward of 40 p was made if two tickets of each colour were picked. Was a good time had byall?In this case, all types of outcome (red, blue <strong>and</strong> green) have the same probabilities. Theprobability of obtaining any given combination of tickets is given by the multinomialdistribution with n =6,k =3<strong>and</strong>p i = 1 , i =1, 2, 3.31208


30.15 IMPORTANT JOINT DISTRIBUTIONS(i) The probability of picking six tickets of the same colour is given byPr (six of the same colour) = 3 × 6! ( ) 6 ( ) 0 ( ) 0 1 1 1= 16!0!0! 3 3 3 243 .The factor of 3 is present because there are three different colours.(ii) The probability of picking five tickets of one colour <strong>and</strong> one ticket of anothercolour isPr(five of one colour; one of another) = 3 × 2 × 6! ( ) 5 ( ) 1 ( ) 0 1 1 1= 4 5!1!0! 3 3 3 81 .The factors of 3 <strong>and</strong> 2 are included because there are three ways to choose thecolour of the five matching tickets, <strong>and</strong> then two ways to choose the colour of theremaining ticket.(iii) Finally, the probability of picking two tickets of each colour isPr (two of each colour) = 6! ( ) 2 ( ) 2 ( ) 2 1 1 1= 102!2!2! 3 3 3 81 .Thus the expected return to any patron was, in pence,( 1100243 + 4 ) (+ 40 × 10 )=10.29.8181A good time was had by all but the stallholder! ◭30.15.2 The multivariate Gaussian distributionA particularly interesting multivariate distribution is provided by the generalisationof the Gaussian distribution to multiple r<strong>and</strong>om variables X i , i =1, 2,...,n.If the expectation value of X i is E(X i )=µ i then the general <strong>for</strong>m of the PDF isgiven by[f(x 1 ,x 2 ,...,x n )=N exp − 1 2∑ ∑ij]a ij (x i − µ i )(x j − µ j ) ,where a ij = a ji <strong>and</strong> N is a normalisation constant that we give below. If we writethe column vectors x =(x 1 x 2 ··· x n ) T <strong>and</strong> µ =(µ 1 µ 2 ··· µ n ) T ,<strong>and</strong>denote the matrix with elements a ij by A thenf(x) =f(x 1 ,x 2 ,...,x n )=N exp [ − 1 2 (x − µ)T A(x − µ) ] ,where A is symmetric. Using the same method as that used to derive (30.145) itis straight<strong>for</strong>ward to show that the MGF of f(x) is given byM(t 1 ,t 2 ,...,t n )=exp ( µ T t + 1 2 tT A −1 t ) ,where the column matrix t =(t 1 t 2 ··· t n ) T . From the MGF, we find thatE[X i X j ]= ∂2 M(0, 0,...,0)= µ i µ j +(A −1 ) ij ,∂t i ∂t j1209


PROBABILITY<strong>and</strong> thus, using (30.135), we obtainCov[X i ,X j ]=E[(X i − µ i )(X j − µ j )] = (A −1 ) ij .Hence A is equal to the inverse of the covariance matrix V of the X i , see (30.139).Thus, with the correct normalisation, f(x) is given byf(x) =1(2π) n/2 (det V) 1/2 exp [ − 1 2 (x − µ)T V −1 (x − µ) ] .(30.148)◮Evaluate the integral∫I = exp [ − 1 (x − 2 µ)T V −1 (x − µ) ] d n x,∞where V is a symmetric matrix, <strong>and</strong> hence verify the normalisation in (30.148).We begin by making the substitution y = x − µ to obtain∫I = exp(− 1 2 yT V −1 y) d n y.∞Since V is a symmetric matrix, it may be diagonalised by an orthogonal trans<strong>for</strong>mation tothe new set of variables y ′ = S T y,whereS is the orthogonal matrix with the normalisedeigenvectors of V as its columns (see section 8.16). In this new basis, the matrix V becomesV ′ = S T VS =diag(λ 1 ,λ 2 ,...,λ n ),where the λ i are the eigenvalues of V. Also, since S is orthogonal, det S = ±1, <strong>and</strong> sod n y = |det S| d n y ′ = d n y ′ .Thus we can write I as∫ ∞ ∫ ∞ ∫ ( )∞n∑ y ′ 2iI = ··· exp − dy 1 ′ dy 2 ′ ···dy n′−∞ −∞ −∞2λ ii=1n∏∫ ( )∞= exp − y′ 2idy i ′ =(2π) n/2 (λ 1 λ 2 ···λ n ) 1/2 , (30.149)i=1−∞ 2λ iwhere we have used the st<strong>and</strong>ard integral ∫ ∞−∞ exp(−αy2 ) dy =(π/α) 1/2 (see subsection6.4.2). From section 8.16, however, we note that the product of eigenvalues in (30.149) isequal to det V. Thus we finally obtainI =(2π) n/2 (det V) 1/2 ,<strong>and</strong> hence the normalisation in (30.148) ensures that f(x) integrates to unity. ◭The above example illustrates some importants points concerning the multivariateGaussian distribution. In particular, we note that the Y i′ are independentGaussian variables with mean zero <strong>and</strong> variance λ i . Thus, given a general set ofn Gaussian variables x with means µ <strong>and</strong> covariance matrix V, one can alwaysper<strong>for</strong>m the above trans<strong>for</strong>mation to obtain a new set of variables y ′ , which arelinear combinations of the old ones <strong>and</strong> are distributed as independent Gaussianswith zero mean <strong>and</strong> variances λ i .This result is extremely useful in proving many of the properties of the mul-1210


30.16 EXERCISEStivariate Gaussian. For example, let us consider the quadratic <strong>for</strong>m (multipliedby 2) appearing in the exponent of (30.148) <strong>and</strong> write it as χ 2 n,i.e.χ 2 n =(x − µ) T V −1 (x − µ). (30.150)From (30.149), we see that we may also write it asn∑χ 2 y ′ 2in = ,λ ii=1which is the sum of n independent Gaussian variables with mean zero <strong>and</strong> unitvariance. Thus, as our notation implies, the quantity χ 2 n is distributed as a chisquaredvariable of order n. As illustrated in exercise 30.40, if the variables X i arerequired to satisfy m linear constraints of the <strong>for</strong>m ∑ ni=1 c iX i =0thenχ 2 n definedin (30.150) is distributed as a chi-squared variable of order n − m.30.16 Exercises30.1 By shading or numbering Venn diagrams, determine which of the following arevalid relationships between events. For those that are, prove the relationshipusing de Morgan’s laws.(a) ( ¯X ∪ Y )=X ∩ Ȳ .(b) ¯X ∪ Ȳ = (X ∪ Y ).(c) (X ∪ Y ) ∩ Z =(X ∪ Z) ∩ Y .(d) X ∪ (Y ∩ Z) =(X ∪ Ȳ ) ∩ ¯Z.(e) X ∪ (Y ∩ Z) =(X ∪ Ȳ ) ∪ ¯Z.30.2 Given that events X,Y <strong>and</strong> Z satisfy(X ∩ Y ) ∪ (Z ∩ X) ∪ ( ¯X ∪ Ȳ )=(Z ∪ Ȳ ) ∪{[( ¯Z ∪ ¯X) ∪ ( ¯X ∩ Z)] ∩ Y },prove that X ⊃ Y , <strong>and</strong> that either X ∩ Z = ∅ or Y ⊃ Z.30.3 A <strong>and</strong> B each have two unbiased four-faced dice, the four faces being numbered1, 2, 3, 4. Without looking, B tries to guess the sum x of the numbers on thebottom faces of A’s two dice after they have been thrown onto a table. If theguess is correct B receives x 2 euros, but if not he loses x euros.Determine B’s expected gain per throw of A’s dice when he adopts each of thefollowing strategies:(a) he selects x at r<strong>and</strong>om in the range 2 ≤ x ≤ 8;(b) he throws his own two dice <strong>and</strong> guesses x to be whatever they indicate;(c) he takes your advice <strong>and</strong> always chooses the same value <strong>for</strong> x. Which numberwould you advise?30.4 Use the method of induction to prove equation (30.16), the probability additionlaw <strong>for</strong> the union of n general events.30.5 Two duellists, A <strong>and</strong> B, take alternate shots at each other, <strong>and</strong> the duel is overwhen a shot (fatal or otherwise!) hits its target. Each shot fired by A has aprobability α of hitting B, <strong>and</strong> each shot fired by B has a probability β of hittingA. Calculate the probabilities P 1 <strong>and</strong> P 2 , defined as follows, that A will win sucha duel: P 1 , A fires the first shot; P 2 , B fires the first shot.If they agree to fire simultaneously, rather than alternately, what is the probabilityP 3 that A will win, i.e. hit B without being hit himself?1211


PROBABILITY30.6 X 1 ,X 2 ,...,X n are independent, identically distributed, r<strong>and</strong>om variables drawnfrom a uni<strong>for</strong>m distribution on [0, 1]. The r<strong>and</strong>om variables A <strong>and</strong> B are definedbyA =min(X 1 ,X 2 ,...,X n ), B = max(X 1 ,X 2 ,...,X n ).For any fixed k such that 0 ≤ k ≤ 1 , find the probability, p 2 n, that bothA ≤ k <strong>and</strong> B ≥ 1 − k.Check your general <strong>for</strong>mula by considering directly the cases (a) k =0,(b)k = 1 , 2(c) n =1<strong>and</strong>(d)n =2.30.7 A tennis tournament is arranged on a straight knockout basis <strong>for</strong> 2 n players,<strong>and</strong> <strong>for</strong> each round, except the final, opponents <strong>for</strong> those still in the competitionare drawn at r<strong>and</strong>om. The quality of the field is so even that in any match it isequally likely that either player will win. Two of the players have surnames thatbegin with ‘Q’. Find the probabilities that they play each other(a) in the final,(b) at some stage in the tournament.30.8 This exercise shows that the odds are hardly ever ‘evens’ when it comes to dicerolling.(a) Gamblers A <strong>and</strong> B each roll a fair six-faced die, <strong>and</strong> B wins if his score isstrictly greater than A’s. Show that the odds are 7 to 5 in A’s favour.(b) Calculate the probabilities of scoring a total T from two rolls of a fair die<strong>for</strong> T =2, 3,...,12. Gamblers C <strong>and</strong> D each roll a fair die twice <strong>and</strong> scorerespective totals T C <strong>and</strong> T D , D winning if T D >T C . Realising that the oddsare not equal, D insists that C should increase her stake <strong>for</strong> each game. Cagrees to stake £1.10 per game, as compared to D’s £1.00 stake. Who willshow a profit?30.9 An electronics assembly firm buys its microchips from three different suppliers;half of them are bought from firm X, whilst firms Y <strong>and</strong> Z supply 30% <strong>and</strong>20%, respectively. The suppliers use different quality-control procedures <strong>and</strong> thepercentages of defective chips are 2%, 4% <strong>and</strong> 4% <strong>for</strong> X, Y <strong>and</strong> Z, respectively.The probabilities that a defective chip will fail two or more assembly-line tests are40%, 60% <strong>and</strong> 80%, respectively, whilst all defective chips have a 10% chanceof escaping detection. An assembler finds a chip that fails only one test. What isthe probability that it came from supplier X?30.10 As every student of probability theory will know, Bayesylvania is awash withnatives, not all of whom can be trusted to tell the truth, <strong>and</strong> lost, <strong>and</strong> apparentlysomewhat deaf, travellers who ask the same question several times in an attemptto get directions to the nearest village.One such traveller finds himself at a T-junction in an area populated by theAsciis <strong>and</strong> Bisciis in the ratio 11 to 5. As is well known, the Biscii always lie, butthe Ascii tell the truth three quarters of the time, giving independent answers toall questions, even to immediately repeated ones.(a) The traveller asks one particular native twice whether he should go to theleft or to the right to reach the local village. Each time he is told ‘left’. Shouldhe take this advice, <strong>and</strong>, if he does, what are his chances of reaching thevillage?(b) The traveller then asks the same native the same question a third time, <strong>and</strong><strong>for</strong> a third time receives the answer ‘left’. What should the traveller do now?Have his chances of finding the village been altered by asking the thirdquestion?1212


30.16 EXERCISES30.11 A boy is selected at r<strong>and</strong>om from amongst the children belonging to families withn children. It is known that he has at least two sisters. Show that the probabilitythat he has k − 1brothersis(n − 1)!(2 n−1 − n)(k − 1)!(n − k)! ,<strong>for</strong> 1 ≤ k ≤ n − 2 <strong>and</strong> zero <strong>for</strong> other values of k. Assume that boys <strong>and</strong> girls areequally likely.30.12 Villages A, B, C <strong>and</strong> D are connected by overhead telephone lines joining AB,AC, BC, BD <strong>and</strong> CD. As a result of severe gales, there is a probability p (thesame <strong>for</strong> each link) that any particular link is broken.(a) Show that the probability that a call can be made from A to B is1 − p 2 − 2p 3 +3p 4 − p 5 .(b) Show that the probability that a call can be made from D to A is1 − 2p 2 − 2p 3 +5p 4 − 2p 5 .30.13 A set of 2N + 1 rods consists of one of each integer length 1, 2,...,2N,2N +1.Three, of lengths a, b <strong>and</strong> c, are selected, of which a is the longest. By consideringthe possible values of b <strong>and</strong> c, determine the number of ways in which a nondegeneratetriangle (i.e. one of non-zero area) can be <strong>for</strong>med (i) if a is even,<strong>and</strong> (ii) if a is odd. Combine these results appropriately to determine the totalnumber of non-degenerate triangles that can be <strong>for</strong>med with the 2N + 1 rods,<strong>and</strong> hence show that the probability that such a triangle can be <strong>for</strong>med from ar<strong>and</strong>om selection (without replacement) of three rods is(N − 1)(4N +1).2(4N 2 − 1)30.14 A certain marksman never misses his target, which consists of a disc of unit radiuswith centre O. The probability that any given shot will hit the target within adistance t of O is t 2 ,<strong>for</strong>0≤ t ≤ 1. The marksman fires n independendent shotsat the target, <strong>and</strong> the r<strong>and</strong>om variable Y is the radius of the smallest circle withcentre O that encloses all the shots. Determine the PDF <strong>for</strong> Y <strong>and</strong> hence findthe expected area of the circle.The shot that is furthest from O is now rejected <strong>and</strong> the corresponding circledetermined <strong>for</strong> the remaining n − 1 shots. Show that its expected area isn − 1n +1 π.30.15 The duration (in minutes) of a telephone call made from a public call-box is ar<strong>and</strong>om variable T . The probability density function of T is⎧⎪⎨ 0 t


PROBABILITYfighting, kittens are removed at r<strong>and</strong>om, one at a time, until peace is restored.Show, by induction, that the expected number of kittens finally remaining isN(x, y) =xy +1 + yx +1 .30.17 If the scores in a cup football match are equal at the end of the normal period ofplay, a ‘penalty shoot-out’ is held in which each side takes up to five shots (fromthe penalty spot) alternately, the shoot-out being stopped if one side acquires anunassailable lead (i.e. has a lead greater than its opponents have shots remaining).If the scores are still level after the shoot-out a ‘sudden death’ competition takesplace.In sudden death each side takes one shot <strong>and</strong> the competition is over if oneside scores <strong>and</strong> the other does not; if both score, or both fail to score, a furthershot is taken by each side, <strong>and</strong> so on. Team 1, which takes the first penalty, hasa probability p 1 , which is independent of the player involved, of scoring <strong>and</strong> aprobability q 1 (= 1 − p 1 ) of missing; p 2 <strong>and</strong> q 2 are defined likewise.Define Pr(i : x, y) as the probability that team i has scored x goals after yattempts, <strong>and</strong> let f(M) be the probability that the shoot-out terminates after atotal of M shots.(a) Prove that the probability that ‘sudden death’ will be needed isf(11+) =5∑( 5 C r ) 2 (p 1 p 2 ) r (q 1 q 2 ) 5−r .r=0(b) Give reasoned arguments (preferably without first looking at the expressionsinvolved) which show thatf(M =2N) =<strong>for</strong> N =3, 4, 5<strong>and</strong>2N−6∑r=0f(M =2N +1)={ }p2 Pr(1 : r, N)Pr(2:5− N + r, N − 1)+ q 2 Pr(1 : 6 − N + r, N)Pr(2:r, N − 1)2N−5∑r=0{ }p1 Pr(1 : 5 − N + r, N)Pr(2:r, N)+ q 1 Pr(1 : r, N)Pr(2:5− N + r, N)<strong>for</strong> N =3, 4.(c) Give an explicit expression <strong>for</strong> Pr(i : x, y) <strong>and</strong> hence show that if the teamsare so well matched that p 1 = p 2 =1/2 thenf(2N) =f(2N +1)=2N−6∑r=02N−5∑r=0( ) 1N!(N − 1)!62 2N r!(N − r)!(6 − N + r)!(2N − 6 − r)! ,( ) 1(N!) 22 2N r!(N − r)!(5 − N + r)!(2N − 5 − r)! .(d) Evaluate these expressions to show that, expressing f(M) in units of 2 −8 ,wehaveM 6 7 8 9 10 11+f(M) 8 24 42 56 63 63Give a simple explanation of why f(10) = f(11+).1214


30.16 EXERCISES30.18 A particle is confined to the one-dimensional space 0 ≤ x ≤ a, <strong>and</strong> classicallyitcanbeinanysmallintervaldx with equal probability. However, quantummechanics gives the result that the probability distribution is proportional tosin 2 (nπx/a), where n is an integer. Find the variance in the particle’s positionin both the classical <strong>and</strong> quantum-mechanical pictures, <strong>and</strong> show that, althoughthey differ, the latter tends to the <strong>for</strong>mer in the limit of large n, in agreementwith the correspondence principle of physics.30.19 A continuous r<strong>and</strong>om variable X has a probability density function f(x); thecorresponding cumulative probability function is F(x). Show that the r<strong>and</strong>omvariable Y = F(X) is uni<strong>for</strong>mly distributed between 0 <strong>and</strong> 1.30.20 For a non-negative integer r<strong>and</strong>om variable X, in addition to the probabilitygenerating function Φ X (t) defined in equation (30.71), it is possible to define theprobability generating function∞∑Ψ X (t) = g n t n ,where g n is the probability that X>n.(a) Prove that Φ X <strong>and</strong> Ψ X are related byΨ X (t) = 1 − Φ X(t).1 − t(b) Show that E[X] is given by Ψ X (1) <strong>and</strong> that the variance of X can beexpressed as 2Ψ ′ X (1) + Ψ X(1) − [Ψ X (1)] 2 .(c) For a particular r<strong>and</strong>om variable X, the probability that X>nis equal toα n+1 ,with0


PROBABILITY30.23 A point P is chosen at r<strong>and</strong>om on the circle x 2 + y 2 = 1. The r<strong>and</strong>om variableX denotes the distance of P from (1, 0). Find the mean <strong>and</strong> variance of X <strong>and</strong>the probability that X is greater than its mean.30.24 As assistant to a celebrated <strong>and</strong> imperious newspaper proprietor, you are giventhe job of running a lottery, in which each of his five million readers will have anequal independent chance, p, of winning a million pounds; you have the job ofchoosing p. However, if nobody wins it will be bad <strong>for</strong> publicity, whilst if morethan two readers do so, the prize cost will more than offset the profit from extracirculation – in either case you will be sacked! Show that, however you choosep, there is more than a 40% chance you will soon be clearing your desk.30.25 The number of errors needing correction on each page of a set of proofs followsa Poisson distribution of mean µ. The cost of the first correction on any page isα <strong>and</strong> that of each subsequent correction on the same page is β. Prove that theaverage cost of correcting a page isα + β(µ − 1) − (α − β)e −µ .30.26 In the game of Blackball, at each turn Muggins draws a ball at r<strong>and</strong>om from abag containing five white balls, three red balls <strong>and</strong> two black balls; after beingrecorded, the ball is replaced in the bag. A white ball earns him $1, whilst ared ball gets him $2; in either case, he also has the option of leaving with hiscurrent winnings or of taking a further turn on the same basis. If he draws ablack ball the game ends <strong>and</strong> he loses all he may have gained previously. Findan expression <strong>for</strong> Muggins’ expected return if he adopts the strategy of drawingup to n balls, provided he has not been eliminated by then.Show that, as the entry fee to play is $3, Muggins should be dissuaded fromplaying Blackball, but, if that cannot be done, what value of n would you advisehim to adopt?30.27 Show that, <strong>for</strong> large r, the value at the maximum of the PDF <strong>for</strong> the gammadistribution of order r with parameter λ is approximately λ/ √ 2π(r − 1).30.28 A husb<strong>and</strong> <strong>and</strong> wife decide that their family will be complete when it includestwo boys <strong>and</strong> two girls – but that this would then be enough! The probabilitythat a new baby will be a girl is p. Ignoring the possibility of identical twins,show that the expected size of their family is( )12pq − 1 − pq ,where q =1− p.30.29 The probability distribution <strong>for</strong> the number of eggs in a clutch is Po(λ), <strong>and</strong> theprobability that each egg will hatch is p (independently of the size of the clutch).Show by direct calculation that the probability distribution <strong>for</strong> the number ofchicks that hatch is Po(λp) <strong>and</strong> so justify the assumptions made in the workedexample at the end of subsection 30.7.1.30.30 A shopper buys 36 items at r<strong>and</strong>om in a supermarket, where, because of thesales tax imposed, the final digit (the number of pence) in the price is uni<strong>for</strong>mly<strong>and</strong> r<strong>and</strong>omly distributed from 0 to 9. Instead of adding up the bill exactly,she rounds each item to the nearest 10 pence, rounding up or down with equalprobability if the price ends in a ‘5’. Should she suspect a mistake if the cashierasks her <strong>for</strong> 23 pence more than she estimated?30.31 Under EU legislation on harmonisation, all kippers are to weigh 0.2000 kg, <strong>and</strong>vendors who sell underweight kippers must be fined by their government. Theweight of a kipper is normally distributed, with a mean of 0.2000 kg <strong>and</strong> ast<strong>and</strong>ard deviation of 0.0100 kg. They are packed in cartons of 100 <strong>and</strong> largequantities of them are sold.Every day, a carton is to be selected at r<strong>and</strong>om from each vendor <strong>and</strong> tested1216


30.16 EXERCISESaccording to one of the following schemes, which have been approved <strong>for</strong> thepurpose.(a) The entire carton is weighed, <strong>and</strong> the vendor is fined 2500 euros if theaverage weight of a kipper is less than 0.1975 kg.(b) Twenty-five kippers are selected at r<strong>and</strong>om from the carton; the vendor isfined 100 euros if the average weight of a kipper is less than 0.1980 kg.(c) Kippers are removed one at a time, at r<strong>and</strong>om, until one has been foundthat weighs more than 0.2000 kg; the vendor is fined 4n(n − 1) euros, wheren is the number of kippers removed.Which scheme should the Chancellor of the Exchequer be urging his governmentto adopt?30.32 In a certain parliament, the government consists of 75 New Socialites <strong>and</strong>the opposition consists of 25 Preservatives. Preservatives never change theirmind, always voting against government policy without a second thought; NewSocialites vote r<strong>and</strong>omly, but with probability p that they will vote <strong>for</strong> their partyleader’s policies.Following a decision by the New Socialites’ leader to drop certain manifestocommitments, N of his party decide to vote consistently with the opposition. Theleader’s advisors reluctantly admit that an election must be called if N is suchthat, at any vote on government policy, the chance of a simple majority in favourwould be less than 80%. Given that p =0.8, estimate the lowest value of N thatwould precipitate an election.30.33 A practical-class demonstrator sends his twelve students to the storeroom tocollect apparatus <strong>for</strong> an experiment, but <strong>for</strong>gets to tell each which type ofcomponent to bring. There are three types, A, B <strong>and</strong> C, held in the stores (inlarge numbers) in the proportions 20%, 30% <strong>and</strong> 50%, respectively, <strong>and</strong> eachstudent picks a component at r<strong>and</strong>om. In order to set up one experiment, oneunit each of A <strong>and</strong> B <strong>and</strong> two units of C are needed. Let Pr(N) be the probabilitythat at least N experiments can be set up.(a) Evaluate Pr(3).(b) Find an expression <strong>for</strong> Pr(N) intermsofk 1 <strong>and</strong> k 2 , the numbers of componentsof types A <strong>and</strong> B respectively selected by the students. Show that Pr(2)canbewritteninthe<strong>for</strong>m6∑ ∑8−iPr(2) = (0.5) 12 12 C i (0.4) i 12−i C j (0.6) j .i=2(c) By considering the conditions under which no experiments can be set up,show that Pr(1) = 0.9145.30.34 The r<strong>and</strong>om variables X <strong>and</strong> Y take integer values, x <strong>and</strong> y, both ≥ 1, <strong>and</strong> suchthat 2x + y ≤ 2a, wherea is an integer greater than 1. The joint probabilitywithin this region is given byPr(X = x, Y = y) =c(2x + y),where c is a constant, <strong>and</strong> it is zero elsewhere.Show that the marginal probability Pr(X = x) is6(a − x)(2x +2a +1)Pr(X = x) = ,a(a − 1)(8a +5)<strong>and</strong> obtain expressions <strong>for</strong> Pr(Y = y), (a) when y is even <strong>and</strong> (b) when y is odd.Show further thatE[Y ]= 6a2 +4a +1.8a +51217j=2


PROBABILITY[ You will need the results about series involving the natural numbers given insubsection 4.2.5. ]30.35 The continuous r<strong>and</strong>om variables X <strong>and</strong> Y have a joint PDF proportional toxy(x − y) 2 with 0 ≤ x ≤ 1 <strong>and</strong> 0 ≤ y ≤ 1. Find the marginal distributions<strong>for</strong> X <strong>and</strong> Y <strong>and</strong> show that they are negatively correlated with correlationcoefficient − 2 . 330.36 A discrete r<strong>and</strong>om variable X takes integer values n =0, 1,... ,N with probabilitiesp n . A second r<strong>and</strong>om variable Y is defined as Y =(X − µ) 2 ,whereµ is theexpectation value of X. Prove that the covariance of X <strong>and</strong> Y is given byN∑N∑Cov[X,Y ]= n 3 p n − 3µ n 2 p n +2µ 3 .n=0Now suppose that X takes all of its possible values with equal probability, <strong>and</strong>hence demonstrate that two r<strong>and</strong>om variables can be uncorrelated, even thoughone is defined in terms of the other.30.37 Two continuous r<strong>and</strong>om variables X <strong>and</strong> Y have a joint probability distributionf(x, y) =A(x 2 + y 2 ),where A is a constant <strong>and</strong> 0 ≤ x ≤ a,0≤ y ≤ a. Show that X <strong>and</strong> Y are negativelycorrelated with correlation coefficient −15/73. By sketching a rough contourmap of f(x, y) <strong>and</strong> marking off the regions of positive <strong>and</strong> negative correlation,convince yourself that this (perhaps counter-intuitive) result is plausible.30.38 A continuous r<strong>and</strong>om variable X is uni<strong>for</strong>mly distributed over the interval [−c, c].Asampleof2n + 1 values of X is selected at r<strong>and</strong>om <strong>and</strong> the r<strong>and</strong>om variableZ is defined as the median of that sample. Show that Z is distributed over [−c, c]with probability density function(2n +1)!f n (z) =(n!) 2 (2c) 2n+1 (c2 − z 2 ) n .Find the variance of Z.30.39 Show that, as the number of trials n becomes large but np i = λ i , i =1, 2,...,k− 1,remains finite, the multinomial probability distribution (30.146),n!M n (x 1 ,x 2 ,...,x k )=x 1 !x 2 ! ···x k ! px 11 px 22 ···px kk ,can be approximated by a multiple Poisson distribution with k − 1 factors:n=0∏k−1M n(x ′ 1 ,x 2 ,...,x k−1 )=i=1e −λ iλ x ii.x i !(Write ∑ k−1ip i = δ <strong>and</strong> express all terms involving subscript k in terms of n <strong>and</strong>δ, either exactly or approximately. You will need to use n! ≈ n ɛ [(n − ɛ)!] <strong>and</strong>(1 − a/n) n ≈ e −a <strong>for</strong> large n.)(a) Verify that the terms of M n ′ when summed over all values of x 1 ,x 2 ,...,x k−1adduptounity.(b) If k =7<strong>and</strong>λ i =9<strong>for</strong>alli =1, 2,...,6, estimate, using the appropriateGaussian approximation, the chance that at least three of x 1 ,x 2 ,...,x 6 willbe 15 or greater.30.40 The variables X i , i =1, 2,...,n, are distributed as a multivariate Gaussian, withmeans µ i <strong>and</strong> a covariance matrix V. IftheX i are required to satisfy the linear1218


30.17 HINTS AND ANSWERSconstraint ∑ ni=1 c iX i =0,wherethec i are constants (<strong>and</strong> not all equal to zero),show that the variableχ 2 n =(x − µ) T V −1 (x − µ)follows a chi-squared distribution of order n − 1.30.17 Hints <strong>and</strong> answers30.1 (a) Yes, (b) no, (c) no, (d) no, (e) yes.30.3 Show that, if p x /16 is the probability that the total will be x, then the corrspondinggain is [p x (x 2 + x) − 16x]/16. (a) A loss of 0.36 euros; (b) a gain of 27/64 euros;(c) a gain of 2.5 euros, provided he takes your advice <strong>and</strong> guesses ‘5’ each time.30.5 P 1 = α(α + β − αβ) −1 ; P 2 = α(1 − β)(α + β − αβ) −1 ; P 3 = P 2 .30.7 If p r is the probability that be<strong>for</strong>e the rth round both players are still in thetournament (<strong>and</strong> there<strong>for</strong>e have not met each other), show thatp r+1 = 1 (2 n+1−r − 214 2 n+1−r − 1 p r <strong>and</strong> hence that p r =2) r−12 n+1−r − 12 n − 1 .(a) The probability that they meet in the final is p n =2 −(n−1) (2 n − 1) −1 .(b) The probability that they meet at some stage in the tournament is given bythe sum ∑ nr=1 p r(2 n+1−r − 1) −1 =2 −(n−1) .30.9 The relative probabilities are X : Y : Z = 50 : 36 : 8 (in units of 10 −4 ); 25/47.30.11 Take A j as the event that a family consists of j boys <strong>and</strong> n − j girls, <strong>and</strong> B asthe event that the boy has at least two sisters. Apply Bayes’ theorem.30.13 (i) For a even, the number of ways is 1 + 3 + 5 + ···+(a − 3), <strong>and</strong> (ii) <strong>for</strong> a oddit is 2+4+6+···+(a − 3). Combine the results <strong>for</strong> a =2m <strong>and</strong> a =2m +1,with m running from 2 to N, to show that the total number of non-degeneratetriangles is given by N(4N +1)(N − 1)/6. The number of possible selections of aset of three rods is (2N + 1)(2N)(2N − 1)/6.30.15 Show that k = e 2 <strong>and</strong> that the average duration of a call is 1 minute. Let p nbe the probability that the call ends during the interval 0.5(n − 1) ≤ t


PROBABILITY30.23 Mean = 4/π. Variance = 2 − (16/π 2 ). Probability that X exceeds its mean=1− (2/π)sin −1 (2/π) =0.561.30.25 Consider, separately, 0, 1 <strong>and</strong> ≥ 2errorsonapage.30.27 Show that the maximum occurs at x =(r − 1)/λ, <strong>and</strong> then use Stirling’s approximationto find the maximum value.30.29 Pr(k chicks hatching) = ∑ ∞n=kPo(n, λ) Bin(n, p).30.31 There is not much to choose between the schemes. In (a) the critical value ofthe st<strong>and</strong>ard variable is −2.5 <strong>and</strong> the average fine would be 15.5 euros. For(b) the corresponding figures are −1.0 <strong>and</strong> 15.9 euros. Scheme (c) is governedby a geometric distribution with p = q = 1 , <strong>and</strong> leads to an expected fine2of ∑ ∞n=1 4n(n − 1)( 1 2 )n . The sum can be evaluated by differentiating the result∑ ∞n=1 pn = p/(1 − p) with respect to p, <strong>and</strong> gives the expected fine as 16 euros.30.33 (a) [12!(0.5) 6 (0.3) 3 (0.2) 3 ]/(6!3!3!)=0.0624.30.35 You will need to establish the normalisation constant <strong>for</strong> the distribution (36),the common mean value (3/5) <strong>and</strong> the common st<strong>and</strong>ard deviation (3/10). Themarginal distributions are f(x) =3x(6x 2 − 8x + 3), <strong>and</strong> the same function of y.The covariance has the value −3/50, yielding a correlation of −2/3.30.37 A = 3/(24a 4 ); µ X = µ Y = 5a/8; σX 2 = σ2 Y = 73a2 /960; E[XY ] = 3a 2 /8;Cov[X,Y ]=−a 2 /64.30.39 (b) With the continuity correction Pr(x i ≥ 15) = 0.0334. The probability that atleast three are 15 or greater is 7.5 × 10 −4 .1220


31StatisticsIn this chapter, we turn to the study of statistics, which is concerned withthe analysis of experimental data. In a book of this nature we cannot hopeto do justice to such a large subject; indeed, many would argue that statisticsbelongs to the realm of experimental science rather than in a mathematicstextbook. Nevertheless, physical scientists <strong>and</strong> engineers are regularly called uponto per<strong>for</strong>m a statistical analysis of their data <strong>and</strong> to present their results in astatistical context. There<strong>for</strong>e, we will concentrate on this aspect of a much moreextensive subject. § 31.1 Experiments, samples <strong>and</strong> populationsWe may regard the product of any experiment as a set of N measurements of somequantity x or set of quantities x,y,...,z. This set of measurements constitutes thedata. Each measurement (or data item) consists accordingly of a single number x ior a set of numbers (x i ,y i ,...,,z i ), where i =1,...,,N. For the moment, we willassume that each data item is a single number, although our discussion can beextended to the more general case.As a result of inaccuracies in the measurement process, or because of intrinsicvariability in the quantity x being measured, one would expect the N measuredvalues x 1 ,x 2 ,...,x N to be different each time the experiment is per<strong>for</strong>med. We may§ There are, in fact, two separate schools of thought concerning statistics: the frequentist approach<strong>and</strong> the Bayesian approach. Indeed, which of these approaches is the more fundamental is still amatter of heated debate. Here we shall concentrate primarily on the more traditional frequentistapproach (despite the preference of some of the authors <strong>for</strong> the Bayesian viewpoint!). For a fullerdiscussion of the frequentist approach one could refer to, <strong>for</strong> example, A. Stuart <strong>and</strong> K. Ord,Kendall’s Advanced Theory of Statistics, vol. 1 (London: Edward Arnold, 1994) or J. F. Kenney<strong>and</strong> E. S. Keeping, Mathematics of Statistics (New York: Van Nostr<strong>and</strong>, 1954). For a discussionof the Bayesian approach one might consult, <strong>for</strong> example, D. S. Sivia, Data Analysis: A BayesianTutorial (Ox<strong>for</strong>d: Ox<strong>for</strong>d University Press, 1996).1221


STATISTICSthere<strong>for</strong>e consider the x i as a set of N r<strong>and</strong>om variables. In the most general case,these r<strong>and</strong>om variables will be described by some N-dimensional joint probabilitydensity function P (x 1 ,x 2 ,...,x N ). § In other words, an experiment consisting of Nmeasurements is considered as a single r<strong>and</strong>om sample from the joint distribution(or population) P (x), where x denotes a point in the N-dimensional data spacehaving coordinates (x 1 ,x 2 ,...,x N ).The situation is simplified considerably if the sample values x i are independent.In this case, the N-dimensional joint distribution P (x) factorises into the productof N one-dimensional distributions,P (x) =P (x 1 )P (x 2 ) ···P (x N ). (31.1)In the general case, each of the one-dimensional distributions P (x i ) may bedifferent. A typical example of this occurs when N independent measurementsare made of some quantity x but the accuracy of the measuring procedure variesbetween measurements.It is often the case, however, that each sample value x i is drawn independentlyfrom the same population. In this case, P (x) is of the <strong>for</strong>m (31.1), but, in addition,P (x i ) has the same <strong>for</strong>m <strong>for</strong> each value of i. The measurements x 1 ,x 2 ,...,x Nare then said to <strong>for</strong>m a r<strong>and</strong>om sample of size N from the one-dimensionalpopulation P (x). This is the most common situation met in practice <strong>and</strong>, unlessstated otherwise, we will assume from now on that this is the case.31.2 Sample statisticsSuppose we have a set of N measurements x 1 ,x 2 ,...,x N . Any function of thesemeasurements (that contains no unknown parameters) is called a sample statistic,or often simply a statistic. Sample statistics provide a means of characterising thedata. Although the resulting characterisation is inevitably incomplete, it is usefulto be able to describe a set of data in terms of a few pertinent numbers. We nowdiscuss the most commonly used sample statistics.§ In this chapter, we will adopt the common convention that P (x) denotes the particular probabilitydensity function that applies to its argument, x. This obviates the need to use a different letter<strong>for</strong> the PDF of each new variable. For example, if X <strong>and</strong> Y are r<strong>and</strong>om variables with differentPDFs, then properly one should denote these distributions by f(x) <strong>and</strong>g(y), say. In our shorth<strong>and</strong>notation, these PDFs are denoted by P (x) <strong>and</strong>P (y), where it is understood that the functional<strong>for</strong>m of the PDF may be different in each case.1222


31.2 SAMPLE STATISTICS188.7 204.7 193.2 169.0168.1 189.8 166.3 200.0Table 31.1 Experimental data giving eight measurements of the round triptime in milliseconds <strong>for</strong> a computer ‘packet’ to travel from Cambridge UK toCambridge MA.31.2.1 AveragesThe simplest number used to characterise a sample is the mean, which<strong>for</strong>Nvalues x i , i =1, 2,...,N, is defined by¯x = 1 NN∑x i . (31.2)In words, the sample mean is the sum of the sample values divided by the numberof values in the sample.◮Table 31.1 gives eight values <strong>for</strong> the round trip time in milliseconds <strong>for</strong> a computer ‘packet’to travel from Cambridge UK to Cambridge MA. Find the sample mean.Using (31.2) the sample mean in milliseconds is given by¯x = 1 (188.7 + 204.7 + 193.2 + 169.0 + 168.1 + 189.8 + 166.3 + 200.0)8= 1479.8 = 184.975.8Since the sample values in table 31.1 are quoted to an accuracy of one decimal place, it isusual to quote the mean to the same accuracy, i.e. as ¯x = 185.0. ◭Strictly speaking the mean given by (31.2) is the arithmetic mean <strong>and</strong> this is byfar the most common definition used <strong>for</strong> a mean. Other definitions of the meanare possible, though less common, <strong>and</strong> include(i) the geometric mean,i=1(ii) the harmonic mean,( N) 1/N∏¯x g = x i , (31.3)¯x h =i=1N∑ Ni=1 1/x , (31.4)i(iii) the root mean square,¯x rms =( ∑Ni=1 x2 iN) 1/2. (31.5)1223


STATISTICSIt should be noted that, ¯x, ¯x h <strong>and</strong> ¯x rms would remain well defined even if somesample values were negative, but the value of ¯x g could then become complex.The geometric mean should not be used in such cases.◮Calculate ¯x g , ¯x h <strong>and</strong> ¯x rms <strong>for</strong> the sample given in table 31.1.The geometric mean is given by (31.3) to be¯x g = (188.7 × 204.7 × ···× 200.0) 1/8 = 184.4.The harmonic mean is given by (31.4) to be8¯x h =(1/188.7) + (1/204.7) + ···+(1/200.0) = 183.9.Finally, the root mean square is given by (31.5) to be¯x rms = [ 18 (188.72 + 204.7 2 + ···+ 200.0 2 ) ] 1/2= 185.5. ◭Two other measures of the ‘average’ of a sample are its mode <strong>and</strong> median. Themode is simply the most commonly occurring value in the sample. A sample maypossess several modes, however, <strong>and</strong> thus it can be misleading in such cases touse the mode as a measure of the average of the sample. The median of a sampleis the halfway point when the sample values x i (i =1, 2,...,N) are arranged inascending (or descending) order. Clearly, this depends on whether the size ofthe sample, N, is odd or even. If N is odd then the median is simply equal tox (N+1)/2 , whereas if N is even the median of the sample is usually taken to be12 (x N/2 + x (N/2)+1 ).◮Find the mode <strong>and</strong> median of the sample given in table 31.1.From the table we see that each sample value occurs exactly once, <strong>and</strong> so any value maybe called the mode of the sample.To find the sample median, we first arrange the sample values in ascending order <strong>and</strong>obtain166.3, 168.1, 169.0, 188.7, 189.8, 193.2, 200.0, 204.7.Since the number of sample values N = 8, which is even, the median of the sample is1(x 2 4 + x 5 )= 1 (188.7 + 189.8) = 189.25. ◭231.2.2 Variance <strong>and</strong> st<strong>and</strong>ard deviationThe variance <strong>and</strong> st<strong>and</strong>ard deviation both give a measure of the spread of valuesin a sample about the sample mean ¯x. Thesample variance is defined bys 2 = 1 NN∑(x i − ¯x) 2 , (31.6)i=11224


31.2 SAMPLE STATISTICS<strong>and</strong> the sample st<strong>and</strong>ard deviation is the positive square root of the samplevariance, i.e.s = √ 1 N∑(x i − ¯x)N2 . (31.7)i=1◮Find the sample variance <strong>and</strong> sample st<strong>and</strong>ard deviation of the data given in table 31.1.We have already found that the sample mean is 185.0 to one decimal place. However,when the mean is to be used in the subsequent calculation of the sample variance it isbetter to use the most accurate value available. In this case the exact value is 184.975, <strong>and</strong>so using (31.6),s 2 = 1 [(188.7 − 184.975) 2 + ···+ (200.0 − 184.975) 2]8= 1608.368= 201.0,where once again we have quoted the result to one decimal place. The sample st<strong>and</strong>arddeviation is then given by s = √ 201.0 =14.2. As it happens, in this case the differencebetween the true mean <strong>and</strong> the rounded value is very small compared with the variationof the individual readings about the mean <strong>and</strong> using the rounded value has a negligibleeffect; however, this would not be so if the difference were comparable to the samplest<strong>and</strong>ard deviation. ◭Using the definition (31.7), it is clear that in order to calculate the st<strong>and</strong>arddeviation of a sample we must first calculate the sample mean. This requirementcan be avoided, however, by using an alternative <strong>for</strong>m <strong>for</strong> s 2 . From (31.6), we seethats 2 = 1 N= 1 NN∑(x i − ¯x) 2i=1N∑x 2 i − 1 Ni=1N∑2x i¯x + 1 Ni=1= x 2 − 2¯x 2 + ¯x 2 = x 2 − ¯x 2N∑¯x 2i=1We may there<strong>for</strong>e write the sample variance s 2 as( ) 2s 2 = x 2 − ¯x 2 = 1 N∑x 2 1N∑i − x i , (31.8)NNi=1from which the sample st<strong>and</strong>ard deviation is found by taking the positive squareroot. Thus, by evaluating the quantities ∑ Ni=1 x i <strong>and</strong> ∑ Ni=1 x2 i <strong>for</strong> our sample, wecan calculate the sample mean <strong>and</strong> sample st<strong>and</strong>ard deviation at the same time.1225i=1


STATISTICS◮Calculate ∑ Ni=1 x i <strong>and</strong> ∑ Ni=1 x2 i <strong>for</strong> the data given in table 31.1 <strong>and</strong> hence find the mean<strong>and</strong> st<strong>and</strong>ard deviation of the sample.From table 31.1, we obtainN∑x i = 188.7 + 204.7+···+ 200.0 = 1479.8,i=1N∑x 2 i = (188.7) 2 + (204.7) 2 + ···+ (200.0) 2 = 275 334.36.i=1Since N = 8, we find as be<strong>for</strong>e (quoting the final results to one decimal place)¯x = 1479.88√( ) 2275 334.36 1479.8= 185.0, s =−=14.2. ◭8831.2.3 Moments <strong>and</strong> central momentsBy analogy with our discussion of probability distributions in section 30.5, thesample mean <strong>and</strong> variance may also be described respectively as the first moment<strong>and</strong> second central moment of the sample. In general, <strong>for</strong> a sample x i , i =1, 2,...,N, we define the rth moment m r <strong>and</strong> rth central moment n r asm r = 1 Nn r = 1 NN∑x r i , (31.9)i=1N∑(x i − m 1 ) r . (31.10)i=1Thus the sample mean ¯x <strong>and</strong> variance s 2 may also be written as m 1 <strong>and</strong> n 2respectively. As is common practice, we have introduced a notation in whicha sample statistic is denoted by the Roman letter corresponding to whicheverGreek letter is used to describe the corresponding population statistic. Thus, weuse m r <strong>and</strong> n r to denote the rth moment <strong>and</strong> central moment of a sample, sincein section 30.5 we denoted the rth moment <strong>and</strong> central moment of a populationby µ r <strong>and</strong> ν r respectively.This notation is particularly useful, since the rth central moment of a sample,m r , may be expressed in terms of the rth- <strong>and</strong> lower-order sample moments n r in away exactly analogous to that derived in subsection 30.5.5 <strong>for</strong> the correspondingpopulation statistics. As discussed in the previous section, the sample variance isgiven by s 2 = x 2 − ¯x 2 but this may also be written as n 2 = m 2 − m 2 1 ,whichistobecompared with the corresponding relation ν 2 = µ 2 −µ 2 1 derived in subsection 30.5.3<strong>for</strong> population statistics. This correspondence also holds <strong>for</strong> higher-order central1226


31.2 SAMPLE STATISTICSmoments of the sample. For example,n 3 = 1 N= 1 NN∑(x i − m 1 ) 3i=1N∑(x 3 i − 3m 1 x 2 i +3m 2 1x i − m 3 1)i=1= m 3 − 3m 1 m 2 +3m 2 1m 1 − m 3 1= m 3 − 3m 1 m 2 +2m 3 1, (31.11)which may be compared with equation (30.53) in the previous chapter.Mirroring our discussion of the normalised central moments γ r of a populationin subsection 30.5.5, we can also describe a sample in terms of the dimensionlessquantitiesg k = n k= n kn k/2 s k ;2g 3 <strong>and</strong> g 4 are called the sample skewness <strong>and</strong> kurtosis. Likewise, it is common todefine the excess kurtosis of a sample by g 4 − 3.31.2.4 Covariance <strong>and</strong> correlationSo far we have assumed that each data item of the sample consists of a singlenumber. Now let us suppose that each item of data consists of a pair of numbers,so that the sample is given by (x i ,y i ), i =1, 2,...,N.We may calculate the sample means, ¯x <strong>and</strong> ȳ, <strong>and</strong> sample variances, s 2 x <strong>and</strong>s 2 y,ofthex i <strong>and</strong> y i values individually but these statistics do not provide anymeasure of the relationship between the x i <strong>and</strong> y i . By analogy with our discussionin subsection 30.12.3 we measure any interdependence between the x i <strong>and</strong> y i interms of the sample covariance, which is given byV xy = 1 N∑(x i − ¯x)(y i − ȳ)Ni=1= (x − ¯x)(y − ȳ)= xy − ¯xȳ. (31.12)Writing out the last expression in full, we obtain the <strong>for</strong>m most useful <strong>for</strong>calculations, which reads(V xy = 1 N) (∑x i y i − 1 N)(∑ N)∑NN 2 x i y i .i=11227i=1i=1


STATISTICSr xy =0.0 r xy =0.1 r xy =0.5yxr xy = −0.7 r xy = −0.9 r xy =0.99Figure 31.1 Scatter plots <strong>for</strong> two-dimensional data samples of size N = 1000,with various values of the correlation r. No scales are plotted, since the valueof r is unaffected by shifts of origin or changes of scale in x <strong>and</strong> y.We may also define the closely related sample correlation byr xy = V xy,s x s ywhich can take values between −1 <strong>and</strong> +1. If the x i <strong>and</strong> y i are independent thenV xy =0=r xy , <strong>and</strong> from (31.12) we see that xy = ¯xȳ. It should also be notedthat the value of r xy is not altered by shifts in the origin or by changes in thescale of the x i or y i . In other words, if x ′ = ax + b <strong>and</strong> y ′ = cy + d, wherea,b, c, d are constants, then r x′ y ′ = r xy. Figure 31.1 shows scatter plots <strong>for</strong> severaltwo-dimensional r<strong>and</strong>om samples x i ,y i of size N = 1000, each with a differentvalue of r xy .◮Ten UK citizens are selected at r<strong>and</strong>om <strong>and</strong> their heights <strong>and</strong> weights are found to be asfollows (to the nearest cm or kg respectively):Person A B C D E F G H I JHeight (cm) 194 168 177 180 171 190 151 169 175 182Weight (kg) 75 53 72 80 75 75 57 67 46 68Calculate the sample correlation between the heights <strong>and</strong> weights.In order to find the sample correlation, we begin by calculating the following sums (wherex i are the heights <strong>and</strong> y i are the weights)∑∑x i = 1757, y i = 668,i1228i


31.3 ESTIMATORS AND SAMPLING DISTRIBUTIONS∑x 2 i = 310 041,i∑yi 2 = 45 746,i∑x i y i = 118 029.ThesampleconsistsofN = 10 pairs of numbers, so the means of the x i <strong>and</strong> of the y i aregiven by ¯x = 175.7 <strong>and</strong>ȳ =66.8. Also, xy = 11 802.9. Similarly, the st<strong>and</strong>ard deviationsof the x i <strong>and</strong> y i are calculated, using (31.8), as√( ) 2310 041 1757s x = − =11.6,s y =10√45 746−1010( ) 2 668=10.6.10Thus the sample correlation is given byxy − ¯xȳ 11 802.9 − (175.7)(66.8)r xy = = =0.54.s x s y (11.6)(10.6)Thus there is a moderate positive correlation between the heights <strong>and</strong> weights of thepeople measured. ◭It is straight<strong>for</strong>ward to generalise the above discussion to data samples ofarbitrary dimension, the only complication being one of notation. We chooseto denote the i th data item from an n-dimensional sample as (x (1)i,x (2)i,...,x (n)i),where the bracketted superscript runs from 1 to n <strong>and</strong> labels the elements withina given data item whereas the subscript i runs from 1 to N <strong>and</strong> labels the dataitems within the sample. In this n-dimensional case, we can define the samplecovariance matrix whose elements areV kl = x (k) x (l) − x (k) x (l)<strong>and</strong> the sample correlation matrix with elementsr kl = V kl.s k s lBoth these matrices are clearly symmetric but are not necessarily positive definite.i31.3 Estimators <strong>and</strong> sampling distributionsIn general, the population P (x) from which a sample x 1 ,x 2 ,...,x N is drawnis unknown. Thecentral aim of statistics is to use the sample values x i to infercertain properties of the unknown population P (x), such as its mean, variance <strong>and</strong>higher moments. To keep our discussion in general terms, let us denote the variousparameters of the population by a 1 ,a 2 ,..., or collectively by a. Moreover, we makethe dependence of the population on the values of these quantities explicit bywriting the population as P (x|a). For the moment, we are assuming that thesample values x i are independent <strong>and</strong> drawn from the same (one-dimensional)population P (x|a), in which caseP (x|a) =P (x 1 |a)P (x 2 |a) ···P (x N |a).1229


STATISTICSSuppose, we wish to estimate the value of one of the quantities a 1 ,a 2 ,...,whichwe will denote simply by a. Since the sample values x i provide our only source ofin<strong>for</strong>mation, any estimate of a must be some function of the x i , i.e. some samplestatistic. Such a statistic is called an estimator of a <strong>and</strong> is usually denoted by â(x),where x denotes the sample elements x 1 ,x 2 ,...,x N .Since an estimator â is a function of the sample values of the r<strong>and</strong>om variablesx 1 ,x 2 ,...,x N , it too must be a r<strong>and</strong>om variable. In other words, if a number ofr<strong>and</strong>om samples, each of the same size N, are taken from the (one-dimensional)population P (x|a) then the value of the estimator â will vary from one sample tothe next <strong>and</strong> in general will not be equal to the true value a. This variation in theestimator is described by its sampling distribution P (â|a). From section 30.14, thisis given byP (â|a) dâ = P (x|a) d N x,where d N x is the infinitesimal ‘volume’ in x-space lying between the ‘surfaces’â(x) =â <strong>and</strong> â(x) =â + dâ. The <strong>for</strong>m of the sampling distribution generallydepends upon the estimator under consideration <strong>and</strong> upon the <strong>for</strong>m of thepopulation from which the sample was drawn, including, as indicated, the truevalues of the quantities a. It is also usually dependent on the sample size N.◮The sample values x 1 ,x 2 ,...,x N are drawn independently from a Gaussian distributionwith mean µ <strong>and</strong> variance σ. Suppose that we choose the sample mean ¯x as our estimatorˆµ of the population mean. Find the sampling distributions of this estimator.The sample mean ¯x is given by¯x = 1 N (x 1 + x 2 + ···+ x N ),where the x i are independent r<strong>and</strong>om variables distributed as x i ∼ N(µ, σ 2 ). From ourdiscussion of multiple Gaussian distributions on page 1189, we see immediately that ¯x willalso be Gaussian distributed as N(µ, σ 2 /N). In other words, the sampling distribution of¯x is given byP (¯x|µ, σ) =[1√2πσ2 /N exp −Note that the variance of this distribution is σ 2 /N. ◭](¯x − µ)2. (31.13)2σ 2 /N31.3.1 Consistency, bias <strong>and</strong> efficiency of estimatorsFor any particular quantity a, we may in fact define any number of differentestimators, each of which will have its own sampling distribution. The qualityof a given estimator â may be assessed by investigating certain properties of itssampling distribution P (â|a). In particular, an estimator â is usually judged onthe three criteria of consistency, bias <strong>and</strong> efficiency, each of which we now discuss.1230


31.3 ESTIMATORS AND SAMPLING DISTRIBUTIONSConsistencyAn estimator â is consistent if its value tends to the true value a in the large-samplelimit, i.e.lim â = a.N→∞Consistency is usually a minimum requirement <strong>for</strong> a useful estimator. An equivalentstatement of consistency is that in the limit of large N the samplingdistribution P (â|a) of the estimator must satisfylim P (â|a) → δ(â − a).N→∞BiasThe expectation value of an estimator â is given by∫∫E[â] = âP (â|a) dâ = â(x)P (x|a) d N x, (31.14)where the second integral extends over all possible values that can be taken bythe sample elements x 1 ,x 2 ,...,x N . This expression gives the expected mean valueof â from an infinite number of samples, each of size N. Thebias of an estimatorâ is then defined asb(a) =E[â] − a. (31.15)We note that the bias b does not depend on the measured sample valuesx 1 ,x 2 ,...,x N . In general, though, it will depend on the sample size N, the functional<strong>for</strong>m of the estimator â <strong>and</strong>, as indicated, on the true properties a ofthe population, including the true value of a itself. If b =0thenâ is called anunbiased estimator of a.◮An estimator â is biased in such a way that E[â] =a + b(a), wherethebiasb(a) is givenby (b 1 − 1)a + b 2 <strong>and</strong> b 1 <strong>and</strong> b 2 are known constants. Construct an unbiased estimator of a.Let us first write E[â] is the clearer <strong>for</strong>mE[â] =a +(b 1 − 1)a + b 2 = b 1 a + b 2 .The task of constructing an unbiased estimator is now trivial, <strong>and</strong> an appropriate choiceis â ′ =(â − b 2 )/b 1 , which (as required) has the expectation valueE[â ′ ]= E[â] − b 2b 1= a. ◭EfficiencyThe variance of an estimator is given by∫∫V [â] = (â − E[â]) 2 P (â|a) dâ = (â(x) − E[â]) 2 P (x|a) d N x(31.16)1231


STATISTICS<strong>and</strong> describes the spread of values â about E[â] that would result from a largenumber of samples, each of size N. An estimator with a smaller variance is saidto be more efficient than one with a larger variance. As we show in the nextsection, <strong>for</strong> any given quantity a of the population there exists a theoretical lowerlimit on the variance of any estimator â. This result is known as Fisher’s inequality(or the Cramér–Rao inequality) <strong>and</strong> readsV [â] ≥(1+ ∂b∂a) 2 /E[ ]− ∂2 ln P∂a 2 , (31.17)where P st<strong>and</strong>s <strong>for</strong> the population P (x|a) <strong>and</strong>b is the bias of the estimator.Denoting the quantity on the RHS of (31.17) by V min ,theefficiency e of anestimator is defined ase = V min /V [â].An estimator <strong>for</strong> which e = 1 is called a minimum-variance or efficient estimator.Otherwise, if e


31.3 ESTIMATORS AND SAMPLING DISTRIBUTIONS<strong>and</strong>, on differentiating twice with respect to µ, we find∂ 2 ln P= − N ∂µ 2 σ . 2This is independent of the x i <strong>and</strong> so its expectation value is also equal to −N/σ 2 .Withbset equal to zero in (31.17), Fisher’s inequality thus states that, <strong>for</strong> any unbiased estimatorˆµ of the population mean,V [ˆµ] ≥ σ2N .Since V [¯x] =σ 2 /N, the sample mean ¯x is a minimum-variance estimator of µ. ◭31.3.2 Fisher’s inequalityAs mentioned above, Fisher’s inequality provides a lower limit on the variance ofany estimator â of the quantity a; itreads(V [â] ≥ 1+ ∂b ) 2 / [ ]E − ∂2 ln P∂a∂a 2 , (31.18)where P st<strong>and</strong>s <strong>for</strong> the population P (x|a) <strong>and</strong>b is the bias of the estimator.We now present a proof of this inequality. Since the derivation is somewhatcomplicated, <strong>and</strong> many of the details are unimportant, this section can be omittedon a first reading. Nevertheless, some aspects of the proof will be useful whenthe efficiency of maximum-likelihood estimators is discussed in section 31.5.◮Prove Fisher’s inequality (31.18).The normalisation of P (x|a) isgivenby∫P (x|a) d N x =1, (31.19)where d N x = dx 1 dx 2 ···dx N <strong>and</strong> the integral extends over all the allowed values of thesample items x i . Differentiating (31.19) with respect to the parameter a, weobtain∫∫∂P ∂ ln P∂a dN x =∂a PdN x =0. (31.20)We note that the second integral is simply the expectation value of ∂ ln P/∂a,wheretheaverage is taken over all possible samples x i , i =1, 2,...,N.Further,byequatingthetwoexpressions <strong>for</strong> ∂E[â]/∂a obtained by differentiating (31.15) <strong>and</strong> (31.14) with respect to awe obtain, dropping the functional dependencies, a second relationship,1+ ∂b ∫∂a = â ∂P ∫∂a dN x =â ∂ ln P∂a PdN x. (31.21)Now, multiplying (31.20) by α(a), where α(a) isany function of a, <strong>and</strong> subtracting theresult from (31.21), we obtain∫[â − α(a)] ∂ ln P∂aPdN x =1+ ∂b∂a .At this point we must invoke the Schwarz inequality proved in subsection 8.1.3. The proof1233


STATISTICSis trivially extended to multiple integrals <strong>and</strong> shows that <strong>for</strong> two real functions, g(x) <strong>and</strong>h(x),(∫ )(∫ ) (∫2g 2 (x) d N x h 2 (x) d N x ≥ g(x)h(x) d x) N . (31.22)If we now let g =[â − α(a)] √ P <strong>and</strong> h =(∂ ln P/∂a) √ P , we find{∫} [ ∫ ( ) 2 (∂ ln P[â − α(a)] 2 Pd N xPd x]N ≥ 1+ ∂b ) 2.∂a∂aOn the LHS, the factor in braces represents the expected spread of â-values around thepoint α(a). The minimum value that this integral may take occurs when α(a) =E[â].Making this substitution, we recognise the integral as the variance V [â],<strong>and</strong>soobtaintheresult(V [â] ≥ 1+ ∂b ) [ 2 ∫ ( ) ] 2 −1∂ ln PPd N x . (31.23)∂a ∂aWe note that the factor in brackets is the expectation value of (∂ ln P/∂a) 2 .Fisher’s inequality is, in fact, often quoted in the <strong>for</strong>m (31.23). We may recover the <strong>for</strong>m(31.18) by noting that on differentiating (31.20) with respect to a we obtain∫ ( ∂ 2 ln PP + ∂ ln P )∂Pd N x =0.∂a 2 ∂a ∂aWriting ∂P/∂a as (∂ ln P/∂a)P <strong>and</strong> rearranging we find that∫ ( ) 2 ∫∂ ln P∂Pd N 2 ln Px = − Pd N x.∂a∂a 2Substituting this result in (31.23) gives(V [â] ≥− 1+ ∂b ) 2 [∫ ∂ 2 ln P∂a ∂a 2] −1Pd N x .Since the factor in brackets is the expectation value of ∂ 2 ln P/∂a 2 , we have recoveredresult (31.18). ◭31.3.3 St<strong>and</strong>ard errors on estimatorsFor a given sample x 1 ,x 2 ,...,x N , we may calculate the value of an estimator â(x)<strong>for</strong> the quantity a. It is also necessary, however, to give some measure of thestatistical uncertainty in this estimate. One way of characterising this uncertaintyis with the st<strong>and</strong>ard deviation of the sampling distribution P (â|a), which is givensimply byσâ =(V [â]) 1/2 . (31.24)If the estimator â(x) were calculated <strong>for</strong> a large number of samples, each of sizeN, then the st<strong>and</strong>ard deviation of the resulting â values would be given by (31.24).Consequently, σâ is called the st<strong>and</strong>ard error on our estimate.In general, however, the st<strong>and</strong>ard error σâ depends on the true values of some1234


31.3 ESTIMATORS AND SAMPLING DISTRIBUTIONSor all of the quantities a <strong>and</strong> they may be unknown. When this occurs, one mustsubstitute estimated values of any unknown quantities into the expression <strong>for</strong> σâin order to obtain an estimated st<strong>and</strong>ard error ˆσâ. One then quotes the result asa = â ± ˆσâ.◮Ten independent sample values x i , i =1, 2,...,10, are drawn at r<strong>and</strong>om from a Gaussi<strong>and</strong>istribution with st<strong>and</strong>ard deviation σ =1. The sample values are as follows (to two decimalplaces):2.22 2.56 1.07 0.24 0.18 0.95 0.73 −0.79 2.09 1.81Estimate the population mean µ, quoting the st<strong>and</strong>ard error on your result.We have shown in the final worked example of subsection 31.3.1 that, in this case, ¯x isa consistent, unbiased, minimum-variance estimator of µ <strong>and</strong> has variance V [¯x] =σ 2 /N.Thus, our estimate of the population mean with its associated st<strong>and</strong>ard error isˆµ = ¯x ± √ σ =1.11 ± 0.32.NIf the true value of σ had not been known, we would have needed to use an estimatedvalue ˆσ in the expression <strong>for</strong> the st<strong>and</strong>ard error. Useful basic estimators of σ are discussedin subsection 31.4.2. ◭It should be noted that the above approach is most meaningful <strong>for</strong> unbiasedestimators. In this case, E[â] =a <strong>and</strong> so σâ describes the spread of â-values aboutthe true value a. For a biased estimator, however, the spread about the true valuea is given by the root mean square error ɛâ, which is defined byɛ 2 â = E[(â − a)2 ]= E[(â − E[â]) 2 ]+(E[â] − a) 2= V [â]+b(a) 2 .We see that ɛ 2 âis the sum of the variance of â <strong>and</strong> the square of the bias <strong>and</strong> socan be interpreted as the sum of squares of statistical <strong>and</strong> systematic errors. Fora biased estimator, it is often more appropriate to quote the result asa = â ± ɛâ.As above, it may be necessary to use estimated values â in the expression <strong>for</strong> theroot mean square error <strong>and</strong> thus to quote only an estimate ˆɛâ of the error.31.3.4 Confidence limits on estimatorsAn alternative (<strong>and</strong> often equivalent) way of quoting a statistical error is with aconfidence interval. Let us assume that, other than the quantity of interest a, thequantities a have known fixed values. Thus we denote the sampling distribution1235


STATISTICSP (â|a)αβâ α (a)â β (a)âFigure 31.2 The sampling distribution P (â|a) ofsomeestimatorâ <strong>for</strong> a givenvalue of a. The shaded regions indicate the two probabilities Pr(â â β (a)) = β.of â by P (â|a). For any particular value of a, one can determine the two valuesâ α (a) <strong>and</strong>â β (a) such thatPr(â â β (a)) =∫ âα(a)−∞∫ ∞â β(a)P (â|a) dâ = α, (31.25)P (â|a) dâ = β. (31.26)This is illustrated in figure 31.2. Thus, <strong>for</strong> any particular value of a, the probabilitythat the estimator â lies within the limits â α (a) <strong>and</strong>â β (a) is given byPr(â α (a) < â


31.3 ESTIMATORS AND SAMPLING DISTRIBUTIONSP (â|a − ) P (â|a + )αâ obsβâFigure 31.3 An illustration of how the observed value of the estimator, â obs ,<strong>and</strong> the given values α <strong>and</strong> β determine the two confidence limits a − <strong>and</strong> a + ,which are such that â α (a + )=â obs = â β (a − ).each of size N, were analysed then the interval [a − ,a + ] would contain the truevalue a on a fraction 1 − α − β of the occasions.The interval [a − ,a + ] is called a confidence interval on a at the confidencelevel 1 − α − β. The values a − <strong>and</strong> a + themselves are called respectively thelower confidence limit <strong>and</strong> the upper confidence limit at this confidence level. Inpractice, the confidence level is often quoted as a percentage. A convenient wayof presenting our results is∫ âobs−∞∫ ∞P (â|a + ) dâ = α, (31.28)P (â|a − ) dâ = β. (31.29)â obsThe confidence limits may then be found by solving these equations <strong>for</strong> a − <strong>and</strong>a + either analytically or numerically. The situation is illustrated graphically infigure 31.3.Occasionally one might not combine the results (31.28) <strong>and</strong> (31.29) but useeither one or the other to provide a one-sided confidence interval on a. Wheneverthe results are combined to provide a two-sided confidence interval, however, theinterval is not specified uniquely by the confidence level 1 − α − β. In other words,there are generally an infinite number of intervals [a − ,a + ] <strong>for</strong> which (31.27) holds.To specify a unique interval, one often chooses α = β, resulting in the centralconfidence interval on a. All cases can be covered by calculating the quantitiesc = â − a − <strong>and</strong> d = a + − â <strong>and</strong> quoting the result of an estimate asa = â +d−c.So far we have assumed that the quantities a other than the quantity of interesta are known in advance. If this is not the case then the construction of confidencelimits is considerably more complicated. This is discussed in subsection 31.3.6.1237


STATISTICS31.3.5 Confidence limits <strong>for</strong> a Gaussian sampling distributionAn important special case occurs when the sampling distribution is Gaussian; ifthe mean is a <strong>and</strong> the st<strong>and</strong>ard deviation is σâ then]1 (â − a)2P (â|a, σâ) = √ exp[−2πσâ2 2σâ2 . (31.30)For almost any (consistent) estimator â, the sampling distribution will tend tothis <strong>for</strong>m in the large-sample limit N →∞, as a consequence of the central limittheorem. For a sampling distribution of the <strong>for</strong>m (31.30), the above procedure<strong>for</strong> determining confidence intervals becomes straight<strong>for</strong>ward. Suppose, from oursample, we obtain the value â obs <strong>for</strong> our estimator. In this case, equations (31.28)<strong>and</strong> (31.29) become)(âobs − a +Φ= α,σâ)(âobs − a −1 − Φ= β,σâwhere Φ(z) is the cumulative probability function <strong>for</strong> the st<strong>and</strong>ard Gaussian distribution,discussed in subsection 30.9.1. Solving these equations <strong>for</strong> a − <strong>and</strong> a + givesa − = â obs − σâΦ −1 (1 − β), (31.31)a + = â obs + σâΦ −1 (1 − α); (31.32)we have used the fact that Φ −1 (α) =−Φ −1 (1−α) to make the equations symmetric.The value of the inverse function Φ −1 (z) can be read off directly from table 30.3,given in subsection 30.9.1. For the normally used central confidence interval onehas α = β. In this case, we see that quoting a result using the st<strong>and</strong>ard error, asa = â ± σâ, (31.33)is equivalent to taking Φ −1 (1 − α) = 1. From table 30.3, we find α =1− 0.8413 =0.1587, <strong>and</strong> so this corresponds to a confidence level of 1 − 2(0.1587) ≈ 0.683.Thus, the st<strong>and</strong>ard error limits give the 68.3% central confidence interval.◮Ten independent sample values x i , i =1, 2,...,10, are drawn at r<strong>and</strong>om from a Gaussi<strong>and</strong>istribution with st<strong>and</strong>ard deviation σ =1. The sample values are as follows (to two decimalplaces):2.22 2.56 1.07 0.24 0.18 0.95 0.73 −0.79 2.09 1.81Find the 90% central confidence interval on the population mean µ.Our estimator ˆµ is the sample mean ¯x. As shown towards the end of section 31.3, thesampling distribution of ¯x is Gaussian with mean E[¯x] <strong>and</strong> variance V [¯x] =σ 2 /N. Sinceσ = 1 in this case, the st<strong>and</strong>ard error is given by σˆx = σ/ √ N =0.32. Moreover, insubsection 31.3.3, we found the mean of the above sample to be ¯x =1.11.1238


31.3 ESTIMATORS AND SAMPLING DISTRIBUTIONSFor the 90% central confidence interval, we require α = β =0.05. From table 30.3, wefindΦ −1 (1 − α) =Φ −1 (0.95) = 1.65,<strong>and</strong> using (31.31) <strong>and</strong> (31.32) we obtaina − = ¯x − 1.65σ¯x =1.11 − (1.65)(0.32) = 0.58,a + = ¯x +1.65σ¯x =1.11 + (1.65)(0.32) = 1.64.Thus, the 90% central confidence interval on µ is [0.58, 1.64]. For comparison, the truevalue used to create the sample was µ =1.◭In the case where the st<strong>and</strong>ard error σâ in (31.33) is not known in advance,one must use a value ˆσâ estimated from the sample. In principle, this complicatessomewhat the construction of confidence intervals, since properly one shouldconsider the two-dimensional joint sampling distribution P (â, ˆσâ|a). Nevertheless,in practice, provided ˆσâ is a fairly good estimate of σâ the above procedure maybe applied with reasonable accuracy. In the special case where the sample valuesx i are drawn from a Gaussian distribution with unknown µ <strong>and</strong> σ, itisinfactpossible to obtain exact confidence intervals on the mean µ, <strong>for</strong> a sample of anysize N, using Student’s t-distribution. This is discussed in subsection 31.7.5.31.3.6 Estimation of several quantities simultaneouslySuppose one uses a sample x 1 ,x 2 ,...,x N to calculate the values of several estimatorsâ 1 , â 2 ,...,â M (collectively denoted by â) of the quantities a 1 ,a 2 ,...,a M(collectively denoted by a) that describe the population from which the sample wasdrawn. The joint sampling distribution of these estimators is an M-dimensionalPDF P (â|a) given byP (â|a) d M â = P (x|a) d N x.◮Sample values x 1 ,x 2 ,...,x N are drawn independently from a Gaussian distribution withmean µ <strong>and</strong> st<strong>and</strong>ard deviation σ. Suppose we choose the sample mean ¯x <strong>and</strong> sample st<strong>and</strong>arddeviation s respectively as estimators ˆµ <strong>and</strong> ˆσ. Find the joint sampling distribution ofthese estimators.Since each data value x i in the sample is assumed to be independent of the others, thejoint probability distribution of sample values is given by[ ∑iP (x|µ, σ) =(2πσ 2 ) −N/2 exp −(x ]i − µ) 2.2σ 2We may rewrite the sum in the exponent as follows:∑(x i − µ) 2 = ∑ (x i − ¯x + ¯x − µ) 2ii= ∑ i(x i − ¯x) 2 +2(¯x − µ) ∑ i= Ns 2 + N(¯x − µ) 2 ,1239(x i − ¯x)+ ∑ i(¯x − µ) 2


STATISTICSwhere in the last line we have used the fact that ∑ i (x i − ¯x) = 0. Hence, <strong>for</strong> given valuesof µ <strong>and</strong> σ, the sampling distribution is in fact a function only of the sample mean ¯x <strong>and</strong>the st<strong>and</strong>ard deviation s. Thus the sampling distribution of ¯x <strong>and</strong> s must satisfy{P (¯x, s|µ, σ) d¯xds=(2πσ 2 ) −N/2 exp − N[(¯x − }µ)2 + s 2 ]dV , (31.34)2σ 2where dV = dx 1 dx 2 ···dx N is an element of volume in the sample space which yieldssimultaneously values of ¯x <strong>and</strong> s that lie within the region bounded by [¯x, ¯x + d¯x] <strong>and</strong>[s, s + ds]. Thus our only remaining task is to express dV in terms of ¯x <strong>and</strong> s <strong>and</strong> theirdifferentials.Let S be the point in sample space representing the sample (x 1 ,x 2 ,...,x N ). For givenvalues of ¯x <strong>and</strong> s, we require the sample values to satisfy both the condition∑x i = N¯x,iwhich defines an (N − 1)-dimensional hyperplane in the sample space, <strong>and</strong> the condition∑(x i − ¯x) 2 = Ns 2 ,iwhich defines an (N − 1)-dimensional hypersphere. Thus S is constrained to lie in theintersection of these two hypersurfaces, which is itself an (N −2)-dimensional hypersphere.Now, the volume of an (N − 2)-dimensional hypersphere is proportional to s N−1 . It follows√that the volume √ dV between two concentric (N − 2)-dimensional hyperspheres of radiusNs <strong>and</strong> N(s + ds) <strong>and</strong>two(N − 1)-dimensional hyperplanes corresponding to ¯x <strong>and</strong>¯x + d¯x isdV = As N−2 ds d¯x,where A is some constant. Thus, substituting this expression <strong>for</strong> dV into (31.34), we find])N(¯x − µ)2P (¯x, s|µ, σ) =C 1 exp[− C2σ 2 2 s N−2 exp(− Ns2 = P (¯x|µ, σ)P (s|σ),2σ 2 (31.35)where C 1 <strong>and</strong> C 2 are constants. We have written P (¯x, s|µ, σ) in this <strong>for</strong>m to show that itseparates naturally into two parts, one depending only on ¯x <strong>and</strong> the other only on s. Thus,¯x <strong>and</strong> s are independent variables. Separate normalisations of the two factors in (31.35)require( ) 1/2 ( ) (N−1)/2 NN1C 1 =<strong>and</strong> C2πσ 2 2 =22σ 2 Γ ( ,1(N − 1))2where the calculation of C 2 requires the use of the gamma function, discussed in theAppendix. ◭The marginal sampling distribution of any one of the estimators â i is givensimply by∫ ∫P (â i |a) = ··· P (â|a) dâ 1 ···dâ i−1 dâ i+1 ···dâ M ,<strong>and</strong> the expectation value E[â i ] <strong>and</strong> variance V [â i ]ofâ i are again given by (31.14)<strong>and</strong> (31.16) respectively. By analogy with the one-dimensional case, the st<strong>and</strong>arderror σâi on the estimator â i is given by the positive square root of V [â i ]. With1240


31.3 ESTIMATORS AND SAMPLING DISTRIBUTIONSseveral estimators, however, it is usual to quote their full covariance matrix. ThisM × M matrix has elements∫V ij =Cov[â i , â j ]= (â i − E[â i ])(â j − E[â j ])P (â|a) d M â∫= (â i − E[â i ])(â j − E[â j ])P (x|a) d N x.Fisher’s inequality can be generalised to the multi-dimensional case. Adaptingthe proof given in subsection 31.3.2, one may show that, in the case where theestimators are efficient <strong>and</strong> have zero bias, the elements of the inverse of thecovariance matrix are given by[ ](V −1 ) ij = E − ∂2 ln P, (31.36)∂a i ∂a jwhere P denotes the population P (x|a) from which the sample is drawn. Thequantity on the RHS of (31.36) is the element F ij of the so-called Fisher matrixF of the estimators.◮Calculate the covariance matrix of the estimators ¯x <strong>and</strong> s in the previous example.As shown in (31.35), the joint sampling distribution P (¯x, s|µ, σ) factorises, <strong>and</strong> so theestimators ¯x <strong>and</strong> s are independent. Thus, we conclude immediately thatCov[¯x, s] =0.Since we have already shown in the worked example at the end of subsection 31.3.1 thatV [¯x] =σ 2 /N, it only remains to calculate V [s]. From (31.35), we find∫ ∞) ( ) r/2E[s r ]=C 2 s N−2+r exp(− Ns2 2 Γ ( 1(N − 1+r))2ds =02σ 2 N Γ ( 1(N − 1)) σr ,2where we have evaluated the integral using the definition of the gamma function given inthe Appendix. Thus, the expectation value of the sample st<strong>and</strong>ard deviation is( ) 1/2(2 Γ1E[s] =N) 2N Γ ( σ, (31.37)1(N − 1))2<strong>and</strong> its variance is given by⎧ [⎨⎩ N − 1 − 2 Γ ( 1N) 2V [s] =E[s 2 ] − (E[s]) 2 = σ2NΓ ( 12 (N − 1)) ] 2⎫⎬⎭We note, in passing, that (31.37) shows that s is a biased estimator of σ. ◭The idea of a confidence interval can also be extended to the case where severalquantities are estimated simultaneously but then the practical construction of aninterval is considerably more complicated. The general approach is to constructan M-dimensional confidence region R in a-space. By analogy with the onedimensionalcase, <strong>for</strong> a given confidence level of (say) 1 − α, one first constructs1241


STATISTICSa region ˆR in â-space, such that∫∫P (â|a) d M â =1− α.ˆRA common choice <strong>for</strong> such a region is that bounded by the ‘surface’ P (â|a) =constant. By considering all possible values a <strong>and</strong> the values of â lying withinthe region ˆR, one can construct a 2M-dimensional region in the combined space(â, a). Suppose now that, from our sample x, the values of the estimators areâ i,obs , i =1, 2,...,M. The intersection of the M ‘hyperplanes’ â i = â i,obs withthe 2M-dimensional region will determine an M-dimensional region which, whenprojected onto a-space, will determine a confidence limit R at the confidencelevel 1 − α. It is usually the case that this confidence region has to be evaluatednumerically.The above procedure is clearly rather complicated in general <strong>and</strong> a simplerapproximate method that uses the likelihood function is discussed in subsection31.5.5. As a consequence of the central limit theorem, however, in thelarge-sample limit, N →∞, the joint sampling distribution P (â|a) will tend, ingeneral, towards the multivariate GaussianP (â|a) =1(2π) M/2 |V| 1/2 exp [ − 1 2 Q(â, a)] , (31.38)where V is the covariance matrix of the estimators <strong>and</strong> the quadratic <strong>for</strong>m Q isgiven byQ(â, a) =(â − a) T V −1 (â − a).Moreover, in the limit of large N, the inverse covariance matrix tends to theFisher matrix F given in (31.36), i.e. V −1 → F.For the Gaussian sampling distribution (31.38), the process of obtaining confidenceintervals is greatly simplified. The surfaces of constant P (â|a) correspondto surfaces of constant Q(â, a), which have the shape of M-dimensional ellipsoidsin â-space, centred on the true values a. In particular, let us suppose that theellipsoid Q(â, a) =c (where c is some constant) contains a fraction 1 − α of thetotal probability. Now suppose that, from our sample x, we obtain the values â obs<strong>for</strong> our estimators. Because of the obvious symmetry of the quadratic <strong>for</strong>m Qwith respect to a <strong>and</strong> â, it is clear that the ellipsoid Q(a, â obs )=c in a-space thatis centred on â obs should contain the true values a with probability 1 − α. ThusQ(a, â obs )=c defines our required confidence region R at this confidence level.This is illustrated in figure 31.4 <strong>for</strong> the two-dimensional case.It remains only to determine the constant c corresponding to the confidencelevel 1 − α. As discussed in subsection 30.15.2, the quantity Q(â, a) is distributedas a χ 2 variable of order M. Thus, the confidence region corresponding to the1242


31.4 SOME BASIC ESTIMATORSâ 2a 2(a)(b)a truea trueâ obsâ obsâ 1a 1Figure 31.4 (a) The ellipse Q(â, a) =c in â-space. (b) The ellipse Q(a, â obs )=cin a-space that corresponds to a confidence region R at the level 1 − α, whenc satisfies (31.39).confidence level 1 − α is given by Q(a, â obs )=c, wheretheconstantc satisfies∫ c0P (χ 2 M) d(χ 2 M)=1− α, (31.39)<strong>and</strong> P (χ 2 M ) is the chi-squared PDF of order M, discussed in subsection 30.9.4. Thisintegral may be evaluated numerically to determine the constant c. Alternatively,some reference books tabulate the values of c corresponding to given confidencelevels <strong>and</strong> various values of M. A representative selection of values of c is givenin table 31.2; there the number of degrees of freedom is denoted by the moreusual n, rather than M.31.4 Some basic estimatorsIn many cases, one does not know the functional <strong>for</strong>m of the population fromwhich a sample is drawn. Nevertheless, in a case where the sample valuesx 1 ,x 2 ,...,x N are each drawn independently from a one-dimensional populationP (x), it is possible to construct some basic estimators <strong>for</strong> the moments <strong>and</strong> centralmoments of P (x). In this section, we investigate the estimating properties of thecommon sample statistics presented in section 31.2. In fact, expectation values<strong>and</strong> variances of these sample statistics can be calculated without prior knowledgeof the functional <strong>for</strong>m of the population; they depend only on the sample size N<strong>and</strong> certain moments <strong>and</strong> central moments of P (x).31.4.1 Population mean µLet us suppose that the parent population P (x) has mean µ <strong>and</strong> variance σ 2 .Anobvious estimator ˆµ of the population mean is the sample mean ¯x. Provided µ<strong>and</strong> σ 2 are both finite, we may apply the central limit theorem directly to obtain1243


STATISTICS% 99 95 10 5 2.5 1 0.5 0.1n =1 1.57 10 −4 3.93 10 −3 2.71 3.84 5.02 6.63 7.88 10.832 2.01 10 −2 0.103 4.61 5.99 7.38 9.21 10.60 13.813 0.115 0.352 6.25 7.81 9.35 11.34 12.84 16.274 0.297 0.711 7.78 9.49 11.14 13.28 14.86 18.475 0.554 1.15 9.24 11.07 12.83 15.09 16.75 20.526 0.872 1.64 10.64 12.59 14.45 16.81 18.55 22.467 1.24 2.17 12.02 14.07 16.01 18.48 20.28 24.328 1.65 2.73 13.36 15.51 17.53 20.09 21.95 26.129 2.09 3.33 14.68 16.92 19.02 21.67 23.59 27.8810 2.56 3.94 15.99 18.31 20.48 23.21 25.19 29.5911 3.05 4.57 17.28 19.68 21.92 24.73 26.76 31.2612 3.57 5.23 18.55 21.03 23.34 26.22 28.30 32.9113 4.11 5.89 19.81 22.36 24.74 27.69 29.82 34.5314 4.66 6.57 21.06 23.68 26.12 29.14 31.32 36.1215 5.23 7.26 22.31 25.00 27.49 30.58 32.80 37.7016 5.81 7.96 23.54 26.30 28.85 32.00 34.27 39.2517 6.41 8.67 24.77 27.59 30.19 33.41 35.72 40.7918 7.01 9.39 25.99 28.87 31.53 34.81 37.16 42.3119 7.63 10.12 27.20 30.14 32.85 36.19 38.58 43.8220 8.26 10.85 28.41 31.41 34.17 37.57 40.00 45.3121 8.90 11.59 29.62 32.67 35.48 38.93 41.40 46.8022 9.54 12.34 30.81 33.92 36.78 40.29 42.80 48.2723 10.20 13.09 32.01 35.17 38.08 41.64 44.18 49.7324 10.86 13.85 33.20 36.42 39.36 42.98 45.56 51.1825 11.52 14.61 34.38 37.65 40.65 44.31 46.93 52.6230 14.95 18.49 40.26 43.77 46.98 50.89 53.67 59.7040 22.16 26.51 51.81 55.76 59.34 63.69 66.77 73.4050 29.71 34.76 63.17 67.50 71.42 76.15 79.49 86.6660 37.48 43.19 74.40 79.08 83.30 88.38 91.95 99.6170 45.44 51.74 85.53 90.53 95.02 100.4 104.2 112.380 53.54 60.39 96.58 101.9 106.6 112.3 116.3 124.890 61.75 69.13 107.6 113.1 118.1 124.1 128.3 137.2100 70.06 77.93 118.5 124.3 129.6 135.8 140.2 149.4Table 31.2 The tabulated values are those which a variable distributed as χ 2with n degrees of freedom exceeds with the given percentage probability. Forexample, a variable having a χ 2 distribution with 14 degrees of freedom takesvalues in excess of 21.06 on 10% of occasions.1244


31.4 SOME BASIC ESTIMATORSexact expressions, valid <strong>for</strong> samples of any size N, <strong>for</strong> the expectation value <strong>and</strong>variance of ¯x. From parts (i) <strong>and</strong> (ii) of the central limit theorem, discussed insection 30.10, we immediately obtainE[¯x] =µ, V [¯x] = σ2N . (31.40)Thus we see that ¯x is an unbiased estimator of µ. Moreover, we note that thest<strong>and</strong>ard error in ¯x is σ/ √ N, <strong>and</strong> so the sampling distribution of ¯x becomes moretightly centred around µ as the sample size N increases. Indeed, since V [¯x] → 0as N →∞, ¯x is also a consistent estimator of µ.In the limit of large N, we may in fact obtain an approximate <strong>for</strong>m <strong>for</strong> thefull sampling distribution of ¯x. Part (iii) of the central limit theorem (see section30.10) tells us immediately that, <strong>for</strong> large N, the sampling distribution of ¯x isgiven approximately by the Gaussian <strong>for</strong>m[ ]1P (¯x|µ, σ) ≈ √2πσ2 /N exp (¯x − µ)2−2σ 2 ./NNote that this does not depend on the <strong>for</strong>m of the original parent population.If, however, the parent population is in fact Gaussian then this result is exact<strong>for</strong> samples of any size N (as is immediately apparent from our discussion ofmultiple Gaussian distributions in subsection 30.9.1).31.4.2 Population variance σ 2An estimator <strong>for</strong> the population variance σ 2 is not so straight<strong>for</strong>ward to defineas one <strong>for</strong> the mean. Complications arise because, in many cases, the true meanof the population µ is not known. Nevertheless, let us begin by considering thecase where in fact µ is known. In this event, a useful estimator is( )̂σ 2 = 1 N∑(x i − µ) 2 1N∑= x 2 i − µ 2 . (31.41)NNi=1i=1◮Show that ̂σ 2 is an unbiased <strong>and</strong> consistent estimator of the population variance σ 2 .The expectation value of ̂σ 2 is given by[E[ ̂σ 2 ]= 1 N]N E ∑− µ 2 = E[x 2 i ] − µ 2 = µ 2 − µ 2 = σ 2 ,i=1x 2 ifrom which we see that the estimator is unbiased. The variance of the estimator is[V [ ̂σ 2 ]=1 N]N V ∑+ V [µ 2 ]= 1 2 N V [x2 i ]= 1 N (µ 4 − µ 2 2),i=1x 2 iin which we have used that fact that V [µ 2 ]=0<strong>and</strong>V [x 2 i ]=E[x4 i ] − (E[x2 i ])2 = µ 4 − µ 2 2 ,1245


STATISTICSwhere µ r is the rth population moment. Since ̂σ 2 is unbiased <strong>and</strong> V [ ̂σ 2 ] → 0asN →∞,showing that it is also a consistent estimator of σ 2 , the result is established. ◭If the true mean of the population is unknown, however, a natural alternativeis to replace µ by ¯x in (31.41), so that our estimator is simply the sample variances 2 given by( ) 2s 2 = 1 N∑x 2 1N∑i − x i .NNi=1In order to determine the properties of this estimator, we must calculate E[s 2 ]<strong>and</strong> V [s 2 ]. This task is straight<strong>for</strong>ward but lengthy. However, <strong>for</strong> the investigationof the properties of a central moment of the sample, there exists a useful trickthat simplifies the calculation. We can assume, with no loss of generality, thatthe mean µ 1 of the population from which the sample is drawn is equal to zero.With this assumption, the population central moments, ν r ,areidenticaltothecorresponding moments µ r , <strong>and</strong> we may per<strong>for</strong>m our calculation in terms of thelatter. At the end, however, we replace µ r by ν r in the final result <strong>and</strong> so obtaina general expression that is valid even in cases where µ 1 ≠0.◮Calculate E[s 2 ] <strong>and</strong> V [s 2 ] <strong>for</strong> a sample of size N.The expectation value of the sample variance s 2 <strong>for</strong> a sample of size N is given by[ ] ⎡( ) ⎤E[s 2 ]= 1 ∑N E x 2 i − 1 2∑N E ⎣ x 2 i⎦ii⎡= 1 N NE[x2 i ] − 1 N E ⎢∑2 ⎣ii=1x 2 i + ∑ i,jj≠ix i x j⎤⎥⎦ . (31.42)The number of terms in the double summation in (31.42) is N(N − 1), so we findE[s 2 ]=E[x 2 i ] − 1 N 2 (NE[x2 i ]+N(N − 1)E[x i x j ]).Now, since the sample elements x i <strong>and</strong> x j are independent, E[x i x j ]=E[x i ]E[x j ]=0,assuming the mean µ 1 of the parent population to be zero. Denoting the rth moment ofthe population by µ r , we thus obtainE[s 2 ]=µ 2 − µ 2N = N − 1N µ 2 = N − 1N σ2 , (31.43)where in the last line we have used the fact that the population mean is zero, <strong>and</strong> soµ 2 = ν 2 = σ 2 . However, the final result is also valid in the case where µ 1 ≠0.Using the above method, we can also find the variance of s 2 , although the algebra israther heavy going. The variance of s 2 is given byV [s 2 ]=E[s 4 ] − (E[s 2 ]) 2 , (31.44)where E[s 2 ] is given by (31.43). We there<strong>for</strong>e need only consider how to calculate E[s 4 ],1246


31.4 SOME BASIC ESTIMATORSwhere s 4 is given bys 4 =[ ∑i x2 iN( ∑− i x ) ] 2 2iN= (∑ i x2 i )2− 2 (∑ i x2 i )(∑ i x i) 2+ (∑ i x i) 4. (31.45)N 2 N 3 N 4We will consider in turn each of the three terms on the RHS. In the first term, the sum( ∑ i x2 i )2 can be written as( ) 2∑x 2 i = ∑ x 4 i + ∑ x 2 i x 2 j ,iii,jj≠iwhere the first sum contains N terms <strong>and</strong> the second contains N(N − 1) terms. Since thesample elements x i <strong>and</strong> x j are assumed independent, we have E[x 2 i x2 j ]=E[x2 i ]E[x2 j ]=µ2 2 ,<strong>and</strong> so⎡( ) ⎤ 2∑E ⎣ x 2 ⎦i = Nµ 4 + N(N − 1)µ 2 2.iTurning to the second term on the RHS of (31.45),( )( ) 2∑ ∑x 2 i x i = ∑ x 4 i + ∑ x 3 i x j + ∑ x 2 i x 2 j + ∑ x 2 i x j x k .iiii,ji,ji,j,kj≠ij≠ik≠j≠iSince the mean of the population has been assumed to equal zero, the expectation valuesof the second <strong>and</strong> fourth sums on the RHS vanish. The first <strong>and</strong> third sums contain N<strong>and</strong> N(N − 1) terms respectively, <strong>and</strong> so⎡( )( ) ⎤ 2∑ ∑E ⎣ x 2 i x i⎦ = Nµ 4 + N(N − 1)µ 2 2.iiFinally, we consider the third term on the RHS of (31.45), <strong>and</strong> write( ) 4∑x i = ∑ x 4 i + ∑ x 3 i x j + ∑ x 2 i x 2 j + ∑ x 2 i x j x k + ∑ x i x j x k x l .iii,ji,ji,j,ki,j,k,lj≠ij≠ik≠j≠il≠k≠j≠iThe expectation values of the second, fourth <strong>and</strong> fifth sums are zero, <strong>and</strong> the first <strong>and</strong> thirdsums contain N <strong>and</strong> 3N(N − 1) terms respectively (<strong>for</strong> the third sum, there are N(N − 1)/2ways of choosing i <strong>and</strong> j, <strong>and</strong> the multinomial coefficient of x 2 i x 2 j is 4!/(2!2!) = 6). Thus⎡( ) ⎤ 4∑E ⎣ x i⎦ = Nµ 4 +3N(N − 1)µ 2 2.iCollecting together terms, we there<strong>for</strong>e obtainE[s 4 (N − 1)2]= µN 3 4 + (N − 1)(N2 − 2N +3)µ 2N2, (31.46)3which, together with the result (31.43), may be substituted into (31.44) to obtain finallyV [s 2 (N − 1)2 (N − 1)(N − 3)]= µN 3 4 − µ 2N 3 2= N − 1 [(N − 1)νN 3 4 − (N − 3)ν2], 2 (31.47)1247


STATISTICSwhere in the last line we have used again the fact that, since the population mean is zero,µ r = ν r . However, result (31.47) holds even when the population mean is not zero. ◭From (31.43), we see that s 2 is a biased estimator of σ 2 , although the biasbecomes negligible <strong>for</strong> large N. However, it immediately follows that an unbiasedestimator of σ 2 is given simply bŷσ 2 =NN − 1 s2 , (31.48)where the multiplicative factor N/(N − 1) is often called Bessel’s correction. Thusin terms of the sample values x i , i =1, 2,...,N, an unbiased estimator of thepopulation variance σ 2 is given bŷσ 2 = 1N − 1N∑(x i − ¯x) 2 . (31.49)Using (31.47), we find that the variance of the estimator ̂σ 2 is( 2V [ ̂σ N2 ]=V [sN − 1) 2 ]= 1 (ν 4 − N − 3 )N N − 1 ν2 2 ,where ν r is the rth central moment of the parent population. We note that,since E[ ̂σ 2 ]=σ 2 <strong>and</strong> V [ ̂σ 2 ] → 0asN →∞, the statistic ̂σ 2 is also a consistentestimator of the population variance.i=131.4.3 Population st<strong>and</strong>ard deviation σThe st<strong>and</strong>ard deviation σ of a population is defined as the positive square root ofthe population variance σ 2 (as, indeed, our notation suggests). Thus, it is commonpractice to take the positive square root of the variance estimator as our estimator<strong>for</strong> σ. Thus, we takeˆσ =( ̂σ2) 1/2, (31.50)where ̂σ 2 is given by either (31.41) or (31.48), depending on whether the populationmean µ is known or unknown. Because of the square root in the definition ofˆσ, it is not possible in either case to obtain an exact expression <strong>for</strong> E[ ˆσ] <strong>and</strong>V [ ˆσ]. Indeed, although in each case the estimator is the positive square root ofan unbiased estimator of σ 2 ,itisnot itself an unbiased estimator of σ. However,the bias does becomes negligible <strong>for</strong> large N.◮Obtain approximate expressions <strong>for</strong> E[ ˆσ] <strong>and</strong> V [ ˆσ] <strong>for</strong> a sample of size N in the casewhere the population mean µ is unknown.As the population mean is unknown, we use (31.50) <strong>and</strong> (31.48) to write our estimator in1248


31.4 SOME BASIC ESTIMATORSthe <strong>for</strong>m( ) 1/2 Nˆσ =s,N − 1where s is the sample st<strong>and</strong>ard deviation. The expectation value of this estimator is givenby( ) 1/2 ( ) 1/2 NNE[ ˆσ] =E[(s 2 ) 1/2 ] ≈(E[s 2 ]) 1/2 = σ.N − 1N − 1An approximate expression <strong>for</strong> the variance of ˆσ may be found using (31.47) <strong>and</strong> is givenbyV [ ˆσ] =NN − 1 V [(s2 ) 1/2 ] ≈NN − 1[ d≈N [ 1N − 1] 2d(s 2 ) (s2 ) 1/2 V [s 2 ]s 2 =E[s 2 ]V [s4s]s 2 ].2 2 =E[s 2 ]Using the expressions (31.43) <strong>and</strong> (31.47) <strong>for</strong> E[s 2 ]<strong>and</strong>V [s 2 ] respectively, we obtainV [ ˆσ] ≈ 1 (ν 4 − N − 3 )4Nν 2 N − 1 ν2 2 . ◭31.4.4 Population moments µ rWe may straight<strong>for</strong>wardly generalise our discussion of estimation of the populationmean µ (= µ 1 ) in subsection 31.4.1 to the estimation of the rth populationmoment µ r . An obvious choice of estimator is the rth sample moment m r .Theexpectation value of m r is given byE[m r ]= 1 NN∑i=1E[x r i ]= Nµ rN = µ r,<strong>and</strong> so it is an unbiased estimator of µ r .The variance of m r may be found in a similar manner, although the calculationis a little more complicated. We find thatV [m r ]=E[(m r − µ r ) 2 ]⎡( ) ⎤ 2= 1 ∑N 2 E ⎣ x r i − Nµ r⎦i⎡⎤= 1 N 2 E ⎣ ∑ x 2ri + ∑ ∑∑x r i x r j − 2Nµ r x r i + N 2 µ 2 ⎦rii j≠ii= 1 N µ 2r − µ 2 r + 1 ∑ ∑N 2 E[x r i x r j]. (31.51)ij≠i1249


STATISTICSHowever, since the sample values x i are assumed to be independent, we haveE[x r i x r j]=E[x r i ]E[x r j]=µ 2 r . (31.52)The number of terms in the sum on the RHS of (31.51) is N(N −1), <strong>and</strong> so we findV [m r ]= 1 N µ 2r − µ 2 r + N − 1Nµ2 r = µ 2r − µ 2 rN . (31.53)Since E[m r ]=µ r <strong>and</strong> V [m r ] → 0asN →∞,therth sample moment m r is alsoa consistent estimator of µ r .◮Find the covariance of the sample moments m r <strong>and</strong> m s <strong>for</strong> a sample of size N.We obtain the covariance of the sample moments m r <strong>and</strong> m s in a similar manner to thatused above to obtain the variance of m r . From the definition of covariance, we haveCov[m r ,m s ]=E[(m r − µ r )(m s − µ s )][( )( )]= 1 ∑ ∑N E x r 2 i − Nµ r x s j − Nµ sij= 1 N 2 E ⎡⎣ ∑ ix r+si+ ∑ i⎤∑∑ ∑x r i x s j − Nµ r x s j − Nµ s x r i + N 2 µ r µ s⎦Assuming the x i to be independent, we may again use result (31.52) to obtainj≠iCov[m r ,m s ]= 1 N [Nµ 2 r+s + N(N − 1)µ r µ s − N 2 µ r µ s − N 2 µ s µ r + N 2 µ r µ s ]= 1 N µ r+s + N − 1Nµ rµ s − µ r µ s= µ r+s − µ r µ s.NWe note that by setting r = s, we recover the expression (31.53) <strong>for</strong> V [m r ]. ◭ji31.4.5 Population central moments ν rWe may generalise the discussion of estimators <strong>for</strong> the second central moment ν 2(or equivalently σ 2 ) given in subsection 31.4.2 to the estimation of the rth centralmoment ν r . In particular, we saw in that subsection that our choice of estimator<strong>for</strong> ν 2 depended on whether the population mean µ 1 is known; the same is true<strong>for</strong> the estimation of ν r .Let us first consider the case in which µ 1 is known. From (30.54), we may writeν r asν r = µ r − r C 1 µ r−1 µ 1 + ···+(−1) k r C k µ r−k µ k 1 + ···+(−1) r−1 ( r C r−1 − 1)µ r 1.If µ 1 is known, a suitable estimator is obviouslyˆν r = m r − r C 1 m r−1 µ 1 + ···+(−1) k r C k m r−k µ k 1 + ···+(−1) r−1 ( r C r−1 − 1)µ r 1,where m r is the rth sample moment. Since µ 1 <strong>and</strong> the binomial coefficients are1250


31.4 SOME BASIC ESTIMATORS(known) constants, it is immediately clear that E[ˆν r ]=ν r ,<strong>and</strong>soˆν r is an unbiasedestimator of ν r . It is also possible to obtain an expression <strong>for</strong> V [ˆν r ], though thecalculation is somewhat lengthy.In the case where the population mean µ 1 is not known, the situation is morecomplicated. We saw in subsection 31.4.2 that the second sample moment n 2 (ors 2 )isnot an unbiased estimator of ν 2 (or σ 2 ). Similarly, the rth central moment ofa sample, n r , is not an unbiased estimator of the rth population central momentν r . However, in all cases the bias becomes negligible in the limit of large N.As we also found in the same subsection, there are complications in calculatingthe expectation <strong>and</strong> variance of n 2 ; these complications increase considerably <strong>for</strong>general r. Nevertheless, we have derived already in this chapter exact expressions<strong>for</strong> the expectation value of the first few sample central moments, which are valid<strong>for</strong> samples of any size N. From (31.40), (31.43) <strong>and</strong> (31.46), we findE[n 1 ]=0,E[n 2 ]= N − 1N ν 2, (31.54)E[n 2 2]= N − 1N 3 [(N − 1)ν 4 +(N 2 − 2N +3)ν2].2By similar arguments it can be shown that(N − 1)(N − 2)E[n 3 ]=N 2 ν 3 , (31.55)E[n 4 ]= N − 1N 3 [(N 2 − 3N +3)ν 4 +3(2N − 3)ν2]. 2 (31.56)From (31.54) <strong>and</strong> (31.55), we see that unbiased estimators of ν 2 <strong>and</strong> ν 3 areˆν 2 =NN − 1 n 2, (31.57)ˆν 3 =N 2(N − 1)(N − 2) n 3, (31.58)where (31.57) simply re-establishes our earlier result that ̂σ 2 = Ns 2 /(N − 1) is anunbiased estimator of σ 2 .Un<strong>for</strong>tunately, the pattern that appears to be emerging in (31.57) <strong>and</strong> (31.58)is not continued <strong>for</strong> higher r, as is seen immediately from (31.56). Nevertheless,in the limit of large N, the bias becomes negligible, <strong>and</strong> often one simply takesˆν r = n r . For large N, it may be shown thatE[n r ] ≈ ν rV [n r ] ≈ 1 N (ν 2r − ν 2 r + r 2 ν 2 ν 2 r−1 − 2rν r−1 ν r+1 )Cov[n r ,n s ] ≈ 1 N (ν r+s − ν r ν s + rsν 2 ν r−1 ν s−1 − rν r−1 ν s+1 − sν s−1 ν r+1 )1251


STATISTICS31.4.6 Population covariance Cov[x, y] <strong>and</strong> correlation Corr[x, y]So far we have assumed that each of our N independent samples consists ofa single number x i . Let us now extend our discussion to a situation in whicheach sample consists of two numbers x i ,y i , which we may consider as beingdrawn r<strong>and</strong>omly from a two-dimensional population P (x, y). In particular, wenow consider estimators <strong>for</strong> the population covariance Cov[x, y] <strong>and</strong> <strong>for</strong> thecorrelation Corr[x, y].When µ x <strong>and</strong> µ y are known, an appropriate estimator of the population covarianceisĈov[x, y] =xy − µ x µ y =This estimator is unbiased since(1N)N∑x i y i − µ x µ y . (31.59)[]E[Ĉov[x, y] = 1 N]N E ∑x i y i − µ x µ y = E[x i y i ] − µ x µ y =Cov[x, y].i=1Alternatively, if µ x <strong>and</strong> µ y are unknown, it is natural to replace µ x <strong>and</strong> µ y in(31.59) by the sample means ¯x <strong>and</strong> ȳ respectively, in which case we recover thesample covariance V xy = xy − ¯xȳ discussed in subsection 31.2.4. This estimatoris biased but an unbiased estimator of the population covariance is obtained by<strong>for</strong>mingĈov[x, y] =i=1NN − 1 V xy. (31.60)◮Calculate the expectation value of the sample covariance V xy <strong>for</strong> a sample of size N.The sample covariance is given by( ) ( )( )1 ∑ 1 ∑ 1 ∑V xy = x i y i − x i y j .NN NiijThus its expectation value is given by[ ] [( )(E[V xy ]= 1 ∑N E x i y i − 1 ∑ ∑N E x 2 iiij⎡⎤= E[x i y i ] − 1 N E ⎢∑2 ⎣ x i y i + ∑ ⎥x i y j ⎦ii,jj≠ix j)]1252


31.4 SOME BASIC ESTIMATORSSince the number of terms in the double sum on the RHS is N(N − 1), we haveE[V xy ]=E[x i y i ] − 1 N 2 (NE[x iy i ]+N(N − 1)E[x i y j ])= E[x i y i ] − 1 N (NE[x iy 2 i ]+N(N − 1)E[x i ]E[y j ])= E[x i y i ] − 1 ( ) N − 1E[xi y i ]+(N − 1)µ x µ y = Cov[x, y],NNwhere we have used the fact that, since the samples are independent, E[x i y j ]=E[x i ]E[y j ]. ◭It is possible to obtain expressions <strong>for</strong> the variances of the estimators (31.59)<strong>and</strong> (31.60) but these quantities depend upon higher moments of the populationP (x, y) <strong>and</strong> are extremely lengthy to calculate.Whether the means µ x <strong>and</strong> µ y are known or unknown, an estimator of thepopulation correlation Corr[x, y] is given byy]Ĉorr[x, y] =Ĉov[x, , (31.61)ˆσ x ˆσ ywhere Ĉov[x, y], ˆσ x <strong>and</strong> ˆσ y are the appropriate estimators of the population covariance<strong>and</strong> st<strong>and</strong>ard deviations. Although this estimator is only asymptoticallyunbiased, i.e. <strong>for</strong> large N, it is widely used because of its simplicity. Once againthe variance of the estimator depends on the higher moments of P (x, y) <strong>and</strong>isdifficult to calculate.In the case in which the means µ x <strong>and</strong> µ y are unknown, a suitable (but biased)estimator isV xyĈorr[x, y] =N =NN − 1 s x s y N − 1 r xy, (31.62)where s x <strong>and</strong> s y are the sample st<strong>and</strong>ard deviations of the x i <strong>and</strong> y i respectively<strong>and</strong> r xy is the sample correlation. In the special case when the parent populationP (x, y) is Gaussian, it may be shown that, if ρ = Corr[x, y],E[r xy ]=ρ − ρ(1 − ρ2 )+O(N −2 ),2N(31.63)V [r xy ]= 1 N (1 − ρ2 ) 2 +O(N −2 ), (31.64)from which the expectation value <strong>and</strong> variance of the estimator Ĉorr[x, y] maybe found immediately.We note finally that our discussion may be extended, without significant alteration,to the general case in which each data item consists of n numbersx i ,y i ,...,z i .1253


STATISTICS31.4.7 A worked exampleTo conclude our discussion of basic estimators, we reconsider the set of experimentaldata given in subsection 31.2.4. We carry the analysis as far as calculatingthe st<strong>and</strong>ard errors in the estimated population parameters, including the populationcorrelation.◮Ten UK citizens are selected at r<strong>and</strong>om <strong>and</strong> their heights <strong>and</strong> weights are found to be asfollows (to the nearest cm or kg respectively):Person A B C D E F G H I JHeight (cm) 194 168 177 180 171 190 151 169 175 182Weight (kg) 75 53 72 80 75 75 57 67 46 68Estimate the means, µ x <strong>and</strong> µ y , <strong>and</strong> st<strong>and</strong>ard deviations, σ x <strong>and</strong> σ y , of the two-dimensionaljoint population from which the sample was drawn, quoting the st<strong>and</strong>ard error on the estimatein each case. Estimate also the correlation Corr[x, y] of the population, <strong>and</strong> quote thest<strong>and</strong>ard error on the estimate under the assumption that the population is a multivariateGaussian.In subsection 31.2.4, we calculated various sample statistics <strong>for</strong> these data. In particular,we found that <strong>for</strong> our sample of size N = 10,¯x = 175.7, ȳ =66.8,s x =11.6, s y =10.6, r xy =0.54.Let us begin by estimating the means µ x <strong>and</strong> µ y . As discussed in subsection 31.4.1, thesample mean is an unbiased, consistent estimator of the population mean. Moreover, thest<strong>and</strong>ard error on ¯x (say) is σ x / √ N. In this case, however, we do not know the true valueof σ x <strong>and</strong> we must estimate it using ̂σ x = √ N/(N − 1)s x . Thus, our estimates of µ x <strong>and</strong>µ y , with associated st<strong>and</strong>ard errors, areˆµ x = ¯x ±ˆµ y = ȳ ±s x√N − 1= 175.7 ± 3.9,s y√N − 1=66.8 ± 3.5.We now turn to estimating σ x <strong>and</strong> σ y . As just mentioned, our estimate of σ x (say)is ̂σ x = √ N/(N − 1)s x . Its variance (see the final line of subsection 31.4.3) is givenapproximately byV [ ˆσ] ≈ 1 (ν 4 − N − 3 )4Nν 2 N − 1 ν2 2 .Since we do not know the true values of the population central moments ν 2 <strong>and</strong> ν 4 ,wemust use their estimated values in this expression. We may take ˆν 2 = ̂σ x 2 =(ˆσ) 2 ,whichwehave already calculated. It still remains, however, to estimate ν 4 .Asimpliedneartheendof subsection 31.4.5, it is acceptable to take ˆν 4 = n 4 . Thus <strong>for</strong> the x i <strong>and</strong> y i values, we have(ˆν 4 ) x = 1 N(ˆν 4 ) y = 1 NN∑(x i − ¯x) 4 = 53 411.6i=1N∑(y i − ȳ) 4 = 27 732.5i=11254


31.5 MAXIMUM-LIKELIHOOD METHODSubstituting these values into (31.50), we obtain( ) 1/2 Nˆσ x =s x ± ( ˆV [ ˆσ x ]) 1/2 =12.2 ± 6.7, (31.65)N − 1( ) 1/2 Nˆσ y =s y ± ( ˆV [ ˆσ y ]) 1/2 =11.2 ± 3.6. (31.66)N − 1Finally, we estimate the population correlation Corr[x, y], which we shall denote by ρ.From (31.62), we haveˆρ =NN − 1 r xy =0.60.Under the assumption that the sample was drawn from a two-dimensional Gaussianpopulation P (x, y), the variance of our estimator is given by (31.64). Since we do not knowthe true value of ρ, we must use our estimate ˆρ. Thus, we find that the st<strong>and</strong>ard error ∆ρin our estimate is given approximately by∆ρ ≈ 10 ( ) 1[1 − (0.60) 2 ] 2 =0.05. ◭9 1031.5 Maximum-likelihood methodThe population from which the sample x 1 ,x 2 ,...,x N is drawn is, in general,unknown. In the previous section, we assumed that the sample values were independent<strong>and</strong> drawn from a one-dimensional population P (x), <strong>and</strong> we consideredbasic estimators of the moments <strong>and</strong> central moments of P (x). We did not, however,assume a particular functional <strong>for</strong>m <strong>for</strong> P (x). We now discuss the processof data modelling, in which a specific <strong>for</strong>m is assumed <strong>for</strong> the population.In the most general case, it will not be known whether the sample values areindependent, <strong>and</strong> so let us consider the full joint population P (x), where x is thepoint in the N-dimensional data space with coordinates x 1 ,x 2 ,...,x N .Wethenadopt the hypothesis H that the probability distribution of the sample values hassome particular functional <strong>for</strong>m L(x; a), dependent on the values of some set ofparameters a i , i =1, 2,...,m. Thus, we haveP (x|a,H)=L(x; a),where we make explicit the conditioning on both the assumed functional <strong>for</strong>m <strong>and</strong>on the parameter values. L(x; a) is called the likelihood function. Hypotheses of thistype <strong>for</strong>m the basis of data modelling <strong>and</strong> parameter estimation. One proposes aparticular model <strong>for</strong> the underlying population <strong>and</strong> then attempts to estimate fromthe sample values x 1 ,x 2 ,...,x N the values of the parameters a defining this model.◮A company measures the duration (in minutes) of the N intervals x i , i = 1, 2,...,Nbetween successive telephone calls received by its switchboard. Suppose that the samplevalues x i are drawn independently from the distribution P (x|τ) =(1/τ)exp(−x/τ), whereτis the mean interval between calls. Calculate the likelihood function L(x; τ).Since the sample values are independent <strong>and</strong> drawn from the stated distribution, the1255


STATISTICSL(x; τ)1N =5L(x; τ)1N =100.50.500τ2 4 6 8 10 12 14 16 18 2000τ2 4 6 8 10 12 14 16 18 20L(x; τ)1N =20L(x; τ)1N =500.50.500τ2 4 6 8 10 12 14 16 18 2000τ2 4 6 8 10 12 14 16 18 20Figure 31.5 Examples of the likelihood function (31.67) <strong>for</strong> samples of differentsize N. In each case, the true value of the parameter is τ =4<strong>and</strong>thesample values x i are indicated by the short vertical lines. For the purposesof illustration, in each case the likelihood function is normalised so that itsmaximum value is unity.likelihood is given byL(x; τ) =P (x i |τ)P (x 2 |τ) ···P (x N |τ)= 1 (τ exp − x )1 1(τ τ exp − x )2··· 1 (τ τ exp − x )Nτ= 1 [τ exp − 1 ]N τ (x 1 + x 2 + ···+ x N ) . (31.67)which is to be considered as a function of τ, given that the sample values x i are fixed. ◭The likelihood function (31.67) depends on just a single parameter τ. Plotsofthe likelihood function, considered as a function of τ, are shown in figure 31.5 <strong>for</strong>samples of different size N. The true value of the parameter τ used to generate thesample values was 4. In each case, the sample values x i are indicated by the shortvertical lines. For the purposes of illustration, the likelihood function in eachcase has been scaled so that its maximum value is unity (this is, in fact, commonpractice). We see that when the sample size is small, the likelihood function isvery broad. As N increases, however, the function becomes narrower (its width isinversely proportional to √ N) <strong>and</strong> tends to a Gaussian-like shape, with its peakcentred on 4, the true value of τ. We discuss these properties of the likelihoodfunction in more detail in subsection 31.5.6.1256


31.5 MAXIMUM-LIKELIHOOD METHODL(x; a)(a)L(x; a)(b)âaâaL(x; a)(c)L(x; a)(d)âaâaFigure 31.6 Typical shapes of one-dimensional likelihood functions L(x; a)encountered in practice, when, <strong>for</strong> illustration purposes, it is assumed that theparameter a is restricted to the range zero to infinity. The ML estimator inthe various cases occurs at: (a) the only stationary point; (b) one of severalstationary points; (c) an end-point of the allowed parameter range that is nota stationary point (although stationary points do exist); (d) an end-point ofthe allowed parameter range in which no stationary point exists.31.5.1 The maximum-likelihood estimatorSince the likelihood function L(x; a) gives the probability density associated withany particular set of values of the parameters a, our best estimate â of theseparameters is given by the values of a <strong>for</strong> which L(x; a) is a maximum. This iscalled the maximum-likelihood estimator (or ML estimator).In general, the likelihood function can have a complicated shape when consideredas a function of a, particularly when the dimensionality of the space ofparameters a 1 ,a 2 ,...,a M is large. It may be that the values of some parametersare either known or assumed in advance, in which case the effective dimensionalityof the likelihood function is reduced accordingly. However, even when thelikelihood depends on just a single parameter a (either intrinsically or as theresult of assuming particular values <strong>for</strong> the remaining parameters), its <strong>for</strong>m maybe complicated when the sample size N is small. Frequently occurring shapes ofone-dimensional likelihood functions are illustrated in figure 31.6, where we haveassumed, <strong>for</strong> definiteness, that the allowed range of the parameter a is zero toinfinity. In each case, the ML estimate â is also indicated. Of course, the ‘shape’ ofhigher-dimensional likelihood functions may be considerably more complicated.In many simple cases, however, the likelihood function L(x; a) has a single1257


STATISTICSmaximum that occurs at a stationary point (the likelihood function is then termedunimodal). In this case, the ML estimators of the parameters a i , i =1, 2,...,M,may be found without evaluating the full likelihood function L(x; a). Instead, onesimply solves the M simultaneous equations∂L∂a i∣∣∣∣a=â=0 <strong>for</strong>i =1, 2,...,M. (31.68)Since ln z is a monotonically increasing function of z (<strong>and</strong> there<strong>for</strong>e has thesame stationary points), it is often more convenient, in fact, to maximise thelog-likelihood function, lnL(x; a), with respect to the a i . Thus, one may, as analternative, solve the equations∂ ln L∂a i∣∣∣∣a=â=0 <strong>for</strong>i =1, 2,...,M. (31.69)Clearly, (31.68) <strong>and</strong> (31.69) will lead to the same ML estimates â of the parameters.In either case, it is, of course, prudent to check that the point a = â is a localmaximum.◮Find the ML estimate of the parameter τ in the previous example, in terms of the measuredvalues x i , i =1, 2,...,N.From (31.67), the log-likelihood function in this case is given byln L(x; τ) =N∑i=1( 1lni/τ)τ e−x = −N∑ (i=1ln τ + x iτ). (31.70)Differentiating with respect to the parameter τ <strong>and</strong> setting the result equal to zero, we find∂ ln L∂τ= −N∑( 1τ − x )i=0.τ 2Thus the ML estimate of the parameter τ is given byi=1ˆτ = 1 NN∑x i , (31.71)i=1which is simply the sample mean of the N measured intervals. ◭In the previous example we assumed that the sample values x i were drawnindependently from the same parent distribution. The ML method is more flexiblethan this restriction might seem to imply, <strong>and</strong> it can equally well be applied tothe common case in which the samples x i are independent but each is drawnfrom a different distribution.1258


31.5 MAXIMUM-LIKELIHOOD METHOD◮In an experiment, N independent measurements x i of some quantity are made. Supposethat the r<strong>and</strong>om measurement error on the i th sample value is Gaussian distributed withmean zero <strong>and</strong> known st<strong>and</strong>ard deviation σ i . Calculate the ML estimate of the true valueµ of the quantity being measured.As the measurements are independent, the likelihood factorises:L(x; µ, {σ k })=N∏P (x i |µ, σ i ),where {σ k } denotes collectively the set of known st<strong>and</strong>ard deviations σ 1 ,σ 2 ,...,σ N .Theindividual distributions are given by1P (x i |µ, σ i )= √ exp[− (x ]i − µ) 2.2πσ2i2σi2<strong>and</strong> so the full log-likelihood function is given byln L(x; µ, {σ k })=− 1 N∑[ln(2πσi 2 )+ (x ]i − µ) 2.2σ 2 i=1iDifferentiating this expression with respect to µ <strong>and</strong> setting the result equal to zero, wefindN∂ ln L∂µ = ∑ x i − µ=0,σi2from which we obtain the ML estimatorˆµ =i=1i=1∑ Ni=1 (x i/σ 2 i )∑ Ni=1 (1/σ2 i ) . (31.72)This estimator is commonly used when averaging data with different statistical weightsw i =1/σi 2. We note that when all the variances σ2 i have the same value the estimatorreduces to the sample mean of the data x i . ◭There is, in fact, no requirement in the ML method that the sample valuesbe independent. As an illustration, we shall generalise the above example to acase in which the measurements x i are not all independent. This would occur, <strong>for</strong>example, if these measurements were based at least in part on the same data.◮In an experiment N measurements x i of some quantity are made. Suppose that the r<strong>and</strong>ommeasurement errors on the samples are drawn from a joint Gaussian distribution with meanzero <strong>and</strong> known covariance matrix V. Calculate the ML estimate of the true value µ of thequantity being measured.From (30.148), the likelihood in this case is given by1L(x; µ, V) =(2π) N/2 |V| exp [ − 1 (x − 1/2 2 µ1)T V −1 (x − µ1) ] ,where x is the column matrix with components x 1 ,x 2 ,...,x N <strong>and</strong> 1 is the column matrixwith all components equal to unity. Thus, the log-likelihood function is given by[ln L(x; µ, V) =− 1 2 N ln(2π)+ln|V| +(x − µ1) T V −1 (x − µ1) ] .1259


STATISTICSDifferentiating with respect to µ <strong>and</strong> setting the result equal to zero gives∂ ln L∂µ = 1T V −1 (x − µ1) =0.Thus, the ML estimator is given by∑ˆµ = 1T V −1 x1 T V −1 1 = i,j (V −1 ) ij x j∑i,j (V .−1 ) ijIn the case of uncorrelated errors in measurement, (V −1 ) ij = δ ij /σi 2 <strong>and</strong> our estimatorreduces to that given in (31.72). ◭In all the examples considered so far, the likelihood function has been effectivelyone-dimensional, either instrinsically or under the assumption that the values ofall but one of the parameters are known in advance. As the following exampleinvolving two parameters shows, the application of the ML method to theestimation of several parameters simultaneously is straight<strong>for</strong>ward.◮In an experiment N measurements x i of some quantity are made. Suppose the r<strong>and</strong>om erroron each sample value is drawn independently from a Gaussian distribution of mean zero butunknown st<strong>and</strong>ard deviation σ (which is the same <strong>for</strong> each measurement). Calculate the MLestimates of the true value µ of the quantity being measured <strong>and</strong> the st<strong>and</strong>ard deviation σof the r<strong>and</strong>om errors.In this case the log-likelihood function is given byln L(x; µ, σ) =− 1 N∑[ln(2πσ 2 )+ (x ]i − µ) 2.2σ 2 i=1Taking partial derivatives of ln L with respect to µ <strong>and</strong> σ <strong>and</strong> setting the results equal tozero at the joint estimate ˆµ, ˆσ, weobtainN∑ x i − ˆµ=0, (31.73)ˆσ 2 i=1N∑ (x i − ˆµ) 2 N∑ 1− =0. (31.74)ˆσ 3 ˆσi=1i=1In principle, one should solve these two equations simultaneously <strong>for</strong> ˆµ <strong>and</strong> ˆσ, but in thiscase we notice that the first is solved immediately byˆµ = 1 N∑x i = ¯x,Ni=1where ¯x is the sample mean. Substituting this result into the second equation, we findˆσ = √ 1 N∑(x i − ¯x)N2 = s,i=1where s is the sample st<strong>and</strong>ard deviation. As shown in subsection 31.4.3, s is a biasedestimator of σ. The reason why the ML method may produce a biased estimator isdiscussed in the next subsection. ◭1260


31.5 MAXIMUM-LIKELIHOOD METHOD31.5.2 Trans<strong>for</strong>mation invariance <strong>and</strong> bias of ML estimatorsAn extremely useful property of ML estimators is that they are invariant toparameter trans<strong>for</strong>mations. Suppose that, instead of estimating some parametera of the assumed population, we wish to estimate some function α(a) oftheparameter. The ML estimator ˆα(a) is given by the value assumed by the functionα(a) at the maximum point of the likelihood, which is simply equal to α(â). Thus,we have the very convenient propertyˆα(a) =α(â).We do not have to worry about the distinction between estimating a <strong>and</strong> estimatinga function of a. Thisisnot true, in general, <strong>for</strong> other estimation procedures.◮A company measures the duration (in minutes) of the N intervals x i , i =1, 2,...,N,between successive telephone calls received by its switchboard. Suppose that the samplevalues x i are drawn independently from the distribution P (x|τ) =(1/τ)exp(−x/τ). FindtheML estimate of the parameter λ =1/τ.This is the same problem as the first one considered in subsection 31.5.1. In terms of thenew parameter λ, the log-likelihood function is given byN∑N∑ln L(x; λ) = ln(λe −λx i)= (ln λ − λx i ).i=1Differentiating with respect to λ <strong>and</strong> setting the result equal to zero, we have∂ ln LN∑( 1=i)∂λ λ − x =0.i=1Thus, the ML estimator of the parameter λ is given by( ) −11N∑ˆλ = x i = ¯x −1 . (31.75)Ni=1Referring back to (31.71), we see that, as expected, the ML estimators of λ <strong>and</strong> τ arerelated by ˆλ =1/ˆτ. ◭Although this invariance property is useful it also means that, in general, MLestimators may be biased. In particular, one must be aware of the fact that evenif â is an unbiased ML estimator of a it does not follow that the estimator ˆα(a) isalso unbiased. In the limit of large N, however, the bias of ML estimators alwaystends to zero. As an illustration, it is straight<strong>for</strong>ward to show (see exercise 31.8)that the ML estimators ˆτ <strong>and</strong> ˆλ in the above example have expectation valuesE[ˆτ] =τ <strong>and</strong> E[ˆλ] = N λ. (31.76)N − 1In fact, since ˆτ = ¯x <strong>and</strong> the sample values are independent, the first result followsimmediately from (31.40). Thus, ˆτ is unbiased, but ˆλ =1/ˆτ is biased, albeit thatthe bias tends to zero <strong>for</strong> large N.1261i=1


STATISTICS31.5.3 Efficiency of ML estimatorsWe showed in subsection 31.3.2 that Fisher’s inequality puts a lower limit on thevariance V [â] of any estimator of the parameter a. Under our hypothesis H onp. 1255, the functional <strong>for</strong>m of the population is given by the likelihood function,i.e. P (x|a,H)=L(x; a). Thus, if this hypothesis is correct, we may replace P byL in Fisher’s inequality (31.18), which then readsV [â] ≥(1+ ∂b∂a) 2 /E[ ]− ∂2 ln L∂a 2 ,where b is the bias in the estimator â. We usually denote the RHS by V min .An important property of ML estimators is that if there exists an efficientestimator â eff ,i.e.one<strong>for</strong>whichV [â eff ]=V min ,thenitmust be the ML estimatoror some function thereof. This is easily shown by replacing P by L in the proofof Fisher’s inequality given in subsection 31.3.2. In particular, we note that theequality in (31.22) holds only if h(x) =cg(x), where c is a constant. Thus, if anefficient estimator â eff exists, this is equivalent to dem<strong>and</strong>ing that∂ ln L= c[â eff − α(a)].∂aNow, the ML estimator â ML is given by∂ ln L∂a ∣ =0 ⇒ c[â eff − α(â ML )] = 0,a=âMLwhich, in turn, implies that â eff must be some function of â ML .◮Show that the ML estimator ˆτ given in (31.71) is an efficient estimator of the parameter τ.As shown in (31.70), the log-likelihood function in this case isN∑ (ln L(x; τ) =− ln τ + x )i.τi=1Differentiating twice with respect to τ, we find∂ 2 ln LN∑( 1=∂τ 2 τ − 2x ) ()i= N 1 − 2 N∑x 2 τ 3 τ 2 i , (31.77)τNi=1i=1<strong>and</strong> so the expectation value of this expression is[ ] ∂ 2 ln LE = N (1 − 2 )∂τ 2 τ 2 τ E[x i] = − N τ , 2where we have used the fact that E[x] =τ. Setting b = 0 in (31.18), we thus find that <strong>for</strong>any unbiased estimator of τ,V [ˆτ] ≥ τ2N .From (31.76), we see that the ML estimator ˆτ = ∑ i x i/N is unbiased. Moreover, usingthe fact that V [x] =τ 2 , it follows immediately from (31.40) that V [ˆτ] =τ 2 /N. Thus ˆτ is aminimum-variance estimator of τ. ◭1262


31.5 MAXIMUM-LIKELIHOOD METHOD31.5.4 St<strong>and</strong>ard errors <strong>and</strong> confidence limits on ML estimatorsThe ML method provides a procedure <strong>for</strong> obtaining a particular set of estimatorsâ ML <strong>for</strong> the parameters a of the assumed population P (x|a). As <strong>for</strong> any other setof estimators, the associated st<strong>and</strong>ard errors, covariances <strong>and</strong> confidence intervalscan be found as described in subsections 31.3.3 <strong>and</strong> 31.3.4.◮A company measures the duration (in minutes) of the 10 intervals x i , i =1, 2,...,10,between successive telephone calls made to its switchboard to be as follows:0.43 0.24 3.03 1.93 1.16 8.65 5.33 6.06 5.62 5.22.Supposing that the sample values are drawn independently from the probability distributionP (x|τ) =(1/τ)exp(−x/τ), find the ML estimate of the mean τ <strong>and</strong> quote an estimate ofthe st<strong>and</strong>ard error on your result.As shown in (31.71) the (unbiased) ML estimator ˆτ in this case is simply the sample mean¯x =3.77. Also, as shown in subsection 31.5.3, ˆτ is a minimum-variance estimator withV [ˆτ] =τ 2 /N. Thus, the st<strong>and</strong>ard error in ˆτ is simply√ τ . (31.78)Nσˆτ =Since we do not know the true value of τ, however, we must instead quote an estimate ˆσˆτof the st<strong>and</strong>ard error, obtained by substituting our estimate ˆτ <strong>for</strong> τ in (31.78). Thus, wequote our final result asτ = ˆτ ±ˆτ √N=3.77 ± 1.19. (31.79)For comparison, the true value used to create the sample was τ =4.◭For the particular problem considered in the above example, it is in fact possibleto derive the full sampling distribution of the ML estimator ˆτ using characteristicfunctions, <strong>and</strong> it is given byP (ˆτ|τ) =NN ˆτ N−1 ((N − 1)! τ N exp − Nˆτ ), (31.80)τwhere N is the size of the sample. This function is plotted in figure 31.7 <strong>for</strong> thecase τ = 4 <strong>and</strong> N = 10, which pertains to the above example. Knowledge of theanalytic <strong>for</strong>m of the sampling distribution allows one to place confidence limitson the estimate ˆτ obtained, as discussed in subsection 31.3.4.◮Using the sample values in the above example, obtain the 68% central confidence intervalon the value of τ.For the sample values given, our observed value of the ML estimator is ˆτ obs =3.77. Thus,from (31.28) <strong>and</strong> (31.29), the 68% central confidence interval [τ − ,τ + ] on the value of τ isfound by solving the equations∫ ˆτobs−∞∫ ∞P (ˆτ|τ + ) dˆτ =0.16,ˆτ obsP (ˆτ|τ − ) dˆτ =0.16,1263


STATISTICSP (ˆτ|τ)0.40.30.20.1002 4 6 8 10 12 14ˆτFigure 31.7 The sampling distribution P (ˆτ|τ) <strong>for</strong> the estimator ˆτ <strong>for</strong> the caseτ =4<strong>and</strong>N = 10.where P (ˆτ|τ) is given by (31.80) with N = 10. The above integrals can be evaluatedanalytically but the calculations are rather cumbersome. It is much simpler to evaluatethem by numerical integration, from which we find [τ − ,τ + ]=[2.86, 5.46]. Alternatively,we could quote the estimate <strong>and</strong> its 68% confidence interval asτ =3.77 +1.69−0.91 .Thus we see that the 68% central confidence interval is not symmetric about the estimatedvalue, <strong>and</strong> differs from the st<strong>and</strong>ard error calculated above. This is a result of the (non-Gaussian) shape of the sampling distribution P (ˆτ|τ), apparent in figure 31.7. ◭In many problems, however, it is not possible to derive the full samplingdistribution of an ML estimator â in order to obtain its confidence intervals.Indeed, one may not even be able to obtain an analytic <strong>for</strong>mula <strong>for</strong> its st<strong>and</strong>arderror σâ. This is particularly true when one is estimating several parameter âsimultaneously, since the joint sampling distribution will be, in general, verycomplicated. Nevertheless, as we discuss below, the likelihood function L(x; a)itself can be used very simply to obtain st<strong>and</strong>ard errors <strong>and</strong> confidence intervals.The justification <strong>for</strong> this has its roots in the Bayesian approach to statistics, asopposed to the more traditional frequentist approach we have adopted here. Wenow give a brief discussion of the Bayesian viewpoint on parameter estimation.31.5.5 The Bayesian interpretation of the likelihood functionAs stated at the beginning of section 31.5, the likelihood function L(x; a) isdefined byP (x|a,H)=L(x; a),1264


31.5 MAXIMUM-LIKELIHOOD METHODwhere H denotes our hypothesis of an assumed functional <strong>for</strong>m. Now, usingBayes’ theorem (see subsection 30.2.3), we may writeP (x|a,H)P (a|H)P (a|x,H)= , (31.81)P (x|H)which provides us with an expression <strong>for</strong> the probability distribution P (a|x,H)of the parameters a, given the (fixed) data x <strong>and</strong> our hypothesis H, in terms ofother quantities that we may assign. The various terms in (31.81) have special<strong>for</strong>mal names, as follows.• The quantity P (a|H) ontheRHSistheprior probability, which represents ourstate of knowledge of the parameter values (given the hypothesis H) be<strong>for</strong>e wehave analysed the data.• This probability is modified by the experimental data x through the likelihoodP (x|a,H).• When appropriately normalised by the evidence P (x|H), this yields the posteriorprobability P (a|x,H), which is the quantity of interest.• The posterior encodes all our inferences about the values of the parameters a.Strictly speaking, from a Bayesian viewpoint, this entire function, P (a|x,H), isthe ‘answer’ to a parameter estimation problem.Given a particular hypothesis, the (normalising) evidence factor P (x|H) isunimportant, since it does not depend explicitly upon the parameter values a.Thus, it is often omitted <strong>and</strong> one considers only the proportionality relationP (a|x,H) ∝ P (x|a,H)P (a|H). (31.82)If necessary, the posterior distribution can be normalised empirically, by requiringthat it integrates to unity, i.e. ∫ P (a|x,H) d m a = 1, where the integral extends overall values of the parameters a 1 ,a 2 ,...,a m .The prior P (a|H) in (31.82) should reflect our entire knowledge concerning thevalues of the parameters a, be<strong>for</strong>e the analysis of the current data x. For example,there may be some physical reason to require some or all of the parameters tolie in a given range. If we are largely ignorant of the values of the parameters,we often indicate this by choosing a uni<strong>for</strong>m (or very broad) prior,P (a|H) = constant,in which case the posterior distribution is simply proportional to the likelihood.In this case, we thus haveP (a|x,H) ∝ L(x; a). (31.83)In other words, if we assume a uni<strong>for</strong>m prior then we can identify the posteriordistribution (up to a normalising factor) with L(x; a), considered as a function ofthe parameters a.1265


STATISTICSThus, a Bayesian statistician considers the ML estimates â ML of the parametersto be the values that maximise the posterior P (a|x,H) under the assumption ofa uni<strong>for</strong>m prior. More importantly, however, a Bayesian would not calculate thest<strong>and</strong>ard error or confidence interval on this estimate using the (classical) methodemployed in subsection 31.3.4. Instead, a far more straight<strong>for</strong>ward approach isadopted. Let us assume, <strong>for</strong> the moment, that one is estimating just a singleparameter a. Using (31.83), we may determine the values a − <strong>and</strong> a + such thatPr(a a + |x,H)= L(x; a) da = β.a +where∫it is assumed that the likelihood has been normalised in such a way thatL(x; a) da = 1. Combining these equations gives∫ a+Pr(a − ≤ a


31.5 MAXIMUM-LIKELIHOOD METHODL(x; τ)0.40.30.20.1002 4 6 8 10 12 14τFigure 31.8 The likelihood function L(x; τ) (normalised to unit area) <strong>for</strong> thesample values given in the worked example in subsection 31.5.4 <strong>and</strong> indicatedhere by short vertical lines.68.3% Bayesian central confidence interval. Even when L(x; a) is not Gaussian,however, (31.85) is often used as a measure of the st<strong>and</strong>ard error.◮For the sample data given in subsection 31.5.4, use the likelihood function to estimate thest<strong>and</strong>ard error ˆσˆτ in the ML estimator ˆτ <strong>and</strong> obtain the Bayesian 68% central confidenceinterval on τ.We showed in (31.67) that the likelihood function in this case is given byL(x; τ) = 1 [τ exp − 1 ]N τ (x 1 + x 2 + ···+ x N ) .where x i , i =1, 2,...,N, denotes the sample value <strong>and</strong> N = 10. This likelihood function isplotted in figure 31.8, after normalising (numerically) to unit area. The short vertical linesin the figure indicate the sample values. We see that the likelihood function peaks at theML estimate ˆτ =3.77 that we found in subsection 31.5.4. Also, from (31.77), we have( )∂ 2 ln L= N 21 − 2 N∑x∂τ 2i .τ τNi=1Remembering that ˆτ = ∑ i x i/N, our estimate of the st<strong>and</strong>ard error in ˆτ is( ∣ ) −1/2ˆσˆτ = − ∂2 ln L= √∂τ∣∣∣τ=ˆτ ˆτ =1.19,2 Nwhich is precisely the estimate of the st<strong>and</strong>ard error we obtained in subsection 31.5.4.It should be noted, however, that in general we would not expect the two estimates ofst<strong>and</strong>ard error made by the different methods to be identical.In order to calculate the Bayesian 68% central confidence interval, we must determinethe values a − <strong>and</strong> a + that satisfy (31.84) with α = β =0.16. In this case, the calculationcan be per<strong>for</strong>med analytically but is somewhat tedious. It is trivial, however, to determinea − <strong>and</strong> a + numerically <strong>and</strong> we find the confidence interval to be [3.16, 6.20]. Thus we canquote our result with 68% central confidence limits asτ =3.77 +2.43−0.61 .1267


STATISTICSBy comparing this result with that given towards the end of subsection 31.5.4, we see that,as we might expect, the Bayesian <strong>and</strong> classical confidence intervals differ somewhat. ◭The above discussion is generalised straight<strong>for</strong>wardly to the estimation ofseveral parameters a 1 ,a 2 ,...,a M simultaneously. The elements of the inverse ofthe covariance matrix of the ML estimators can be approximated by∣(V −1 ) ij = − ∂2 ln L ∣∣∣a=â. (31.86)∂a i ∂a jFrom (31.36), we see that (at least <strong>for</strong> unbiased estimators) the expectation valueof (31.86) is equal to the element F ij of the Fisher matrix.The construction of a multi-dimensional Bayesian confidence region is alsostraight<strong>for</strong>ward. For a given confidence level 1 − α (say), it is most commonto construct the confidence region as the M-dimensional region R in a-space,bounded by the ‘surface’ L(x; a) = constant, <strong>for</strong> which∫L(x; a) d M a =1− α,Rwhere it is assumed that L(x; a) is normalised to unit volume. Moreover, wesee from (31.83) that (assuming a uni<strong>for</strong>m prior probability) we may obtain themarginal posterior distribution <strong>for</strong> any parameter a i simply by integrating thelikelihood function L(x; a) over the other parameters:∫ ∫P (a i |x,H)= ··· L(x; a) da 1 ···da i−1 da i+1 ···da M .Here the integral extends over all possible values of the parameters, <strong>and</strong> againis∫it assumed that the likelihood function is normalised in such a way thatL(x; a) d M a = 1. This marginal distribution can then be used as above todetermine Bayesian confidence intervals on each a i separately.◮Ten independent sample values x i , i =1, 2,...,10, are drawn at r<strong>and</strong>om from a Gaussi<strong>and</strong>istribution with unknown mean µ <strong>and</strong> st<strong>and</strong>ard deviation σ. The sample values are asfollows (to two decimal places):2.22 2.56 1.07 0.24 0.18 0.95 0.73 −0.79 2.09 1.81Find the Bayesian 95% central confidence intervals on µ <strong>and</strong> σ separately.The likelihood function in this case is[L(x; µ, σ) =(2πσ 2 ) −N/2 exp − 1N]∑(x2σ 2 i − µ) 2 . (31.87)i=1Assuming uni<strong>for</strong>m priors on µ <strong>and</strong> σ (over their natural ranges of −∞ → ∞ <strong>and</strong> 0 → ∞respectively), we may identify this likelihood function with the posterior probability, as in(31.83). Thus, the marginal posterior distribution on µ is given by∫ [∞1P (µ|x,H) ∝σ exp − 1N]∑(x N 2σ 2 i − µ) 2 dσ.01268i=1


31.5 MAXIMUM-LIKELIHOOD METHODBy substituting σ =1/u (so that dσ = −du/u 2 ) <strong>and</strong> integrating by parts either (N − 2)/2or (N − 3)/2 times, we findP (µ|x,H) ∝ [ N(¯x − µ) 2 + Ns 2] −(N−1)/2,where we have used the fact that ∑ i (x i − µ) 2 = N(¯x − µ) 2 + Ns 2 , ¯x being the sample mean<strong>and</strong> s 2 the sample variance. We may now obtain the 95% central confidence interval byfinding the values µ − <strong>and</strong> µ + <strong>for</strong> which∫ µ−P (µ|x,H) dµ =0.025 <strong>and</strong> P (µ|x,H) dµ =0.025.−∞µ +The normalisation of the posterior distribution <strong>and</strong> the values µ − <strong>and</strong> µ + are easilyobtained by numerical integration. Substituting in the appropriate values N = 10, ¯x =1.11<strong>and</strong> s =1.01, we find the required confidence interval to be [0.29, 1.97].To obtain a confidence interval on σ, we must first obtain the corresponding marginalposterior distribution. From (31.87), again using the fact that ∑ i (x i−µ) 2 = N(¯x−µ) 2 +Ns 2 ,this is given byP (σ|x,H) ∝ 1 ( ) ∫ ∞]σ exp − Ns2N(¯x − µ)2exp[− dµ.N 2σ 2 −∞ 2σ 2Noting that the integral of a one-dimensional Gaussian is proportional to σ, we concludethatP (σ|x,H) ∝ 1 ( )σ exp − Ns2 .N−1 2σ 2The 95% central confidence interval on σ can then be found in an analogous manner tothat on µ, by solving numerically the equations∫ σ−0P (σ|x,H) dσ =0.025<strong>and</strong>We find the required interval to be [0.76, 2.16]. ◭∫ ∞∫ ∞σ +P (σ|x,H) dσ =0.025.31.5.6 Behaviour of ML estimators <strong>for</strong> large NAs mentioned in subsection 31.3.6, in the large-sample limit N →∞, the samplingdistribution of a set of (consistent) estimators â, whether ML or not, will tend,in general, to a multivariate Gaussian centred on the true values a. Thisisadirect consequence of the central limit theorem. Similarly, in the limit N →∞thelikelihood function L(x; a) also tends towards a multivariate Gaussian but onecentred on the ML estimate(s) â. Thus ML estimators are always asymptoticallyconsistent. This limiting process was illustrated <strong>for</strong> the one-dimensional case byfigure 31.5.Thus, as N becomes large, the likelihood function tends to the <strong>for</strong>mwhere Q denotes the quadratic <strong>for</strong>mL(x; a) =L max exp [ − 1 2 Q(a, â)] ,Q(a, â) =(a − â) T V −1 (a − â)1269


STATISTICS<strong>and</strong> the matrix V −1 is given by( ) ∣V−1ij = − ∂2 ln L ∣∣∣a=â.∂a i ∂a jMoreover, in the limit of large N, this matrix tends to the Fisher matrix given in(31.36), i.e. V −1 → F. Hence ML estimators are asymptotically minimum-variance.Comparison of the above results with those in subsection 31.3.6 shows thatthe large-sample limit of the likelihood function L(x; a) has the same <strong>for</strong>m as thelarge-sample limit of the joint estimator sampling distribution P (â|a). The onlydifference is that P (â|a) is centred in â-space on the true values â = a whereasL(x; a) is centred in a-space on the ML estimates a = â. From figure 31.4 <strong>and</strong> itsaccompanying discussion, we there<strong>for</strong>e conclude that, in the large-sample limit,the Bayesian <strong>and</strong> classical confidence limits on the parameters coincide.31.5.7 Extended maximum-likelihood methodIt is sometimes the case that the number of data items N in our sample is itself ar<strong>and</strong>om variable. Such experiments are typically those in which data are collected<strong>for</strong> a certain period of time during which events occur at r<strong>and</strong>om in some way,as opposed to those in which a prearranged number of data items are collected.In particular, let us consider the case where the sample values x 1 ,x 2 ,...,x N aredrawn independently from some distribution P (x|a) <strong>and</strong> the sample size N is ar<strong>and</strong>om variable described by a Poisson distribution with mean λ, i.e.N ∼ Po(λ).The likelihood function in this case is given byL(x,N; λ, a) = λN ∏NN! e−λ P (x i |a), (31.88)i=1<strong>and</strong> is often called the extended likelihood function. The function L(x; λ, a) canbe used as be<strong>for</strong>e to estimate parameter values or obtain confidence intervals.Two distinct cases arise in the use of the extended likelihood function, dependingon whether the Poisson parameter λ is a function of the parameters a or is anindependent parameter.Let us first consider the case in which λ is a function of the parameters a. From(31.88), we can write the extended log-likelihood function asN∑N∑ln L = N ln λ(a) − λ(a)+ ln P (x i |a) =−λ(a)+ ln[λ(a)P (x i |a)].i=1where we have ignored terms not depending on a. The ML estimates â of theparameters can then be found in the usual way, <strong>and</strong> the ML estimate of thePoisson parameter is simply ˆλ = λ(â). The errors on our estimators â will be, ingeneral, smaller than those obtained in the usual likelihood approach, since ourestimate includes in<strong>for</strong>mation from the value of N as well as the sample values x i .1270i=1


31.6 THE METHOD OF LEAST SQUARESThe other possibility is that λ is an independent parameter <strong>and</strong> not a functionof the parameters a. In this case, the extended log-likelihood function isln L = N ln λ − λ +N∑ln P (x i |a), (31.89)where we have omitted terms not depending on λ or a. Differentiating withrespect to λ <strong>and</strong> setting the result equal to zero, we find that the ML estimate ofλ is simplyˆλ = N.By differentiating (31.89) with respect to the parameters a i <strong>and</strong> setting the resultsequal to zero, we obtain the usual ML estimates â i of their values. In this case,however, the errors in our estimates will be larger, in general, than those in thest<strong>and</strong>ard likelihood approach, since they must include the effect of statisticaluncertainty in the parameter λ.i=131.6 The method of least squaresThe method of least squares is, in fact, just a special case of the method ofmaximum likelihood. Nevertheless, it is so widely used as a method of parameterestimation that it has acquired a special name of its own. At the outset, let ussuppose that a data sample consists of a set of pairs (x i ,y i ), i =1, 2,...,N.Forexample, these data might correspond to the temperature y i measured at variouspoints x i along some metal rod.For the moment, we will suppose that the x i are known exactly, whereas thereexists a measurement error (or noise) n i on each of the values y i . Moreover, letus assume that the true value of y at any position x is given by some functiony = f(x; a) that depends on the M unknown parameters a. Theny i = f(x i ; a)+n i .Our aim is to estimate the values of the parameters a from the data sample.Bearing in mind the central limit theorem, let us suppose that the n i are drawnfrom a Gaussian distribution with no systematic bias <strong>and</strong> hence zero mean. In themost general case the measurement errors n i might not be independent but bedescribed by an N-dimensional multivariate Gaussian with non-trivial covariancematrix N, whose elements N ij =Cov[n i ,n j ] we assume to be known. Under theseassumptions it follows from (30.148), that the likelihood function isL(x, y; a) =1(2π) N/2 |N| 1/2 exp [ − 1 2 χ2 (a) ] ,1271


STATISTICSwhere the quantity denoted by χ 2 is given by the quadratic <strong>for</strong>mN∑χ 2 (a) = [y i − f(x i ; a)](N −1 ) ij [y j − f(x j ; a)] = (y − f) T N −1 (y − f).i,j=1(31.90)In the last equality, we have rewritten the expression in matrix notation bydefining the column vector f with elements f i = f(x i ; a). We note that in the(common) special case in which the measurement errors n i are independent, theircovariance matrix takes the diagonal <strong>for</strong>m N = diag(σ1 2,σ2 2 ,...,σ2 N ), where σ i isthe st<strong>and</strong>ard deviation of the measurement error n i . In this case, the expression(31.90) <strong>for</strong> χ 2 reduces toN∑[ ] 2χ 2 yi − f(x i ; a)(a) =.σ ii=1The least-squares (LS) estimators â LS of the parameter values are defined asthose that minimise the value of χ 2 (a); they are usually determined by solvingthe M equations∣∂χ 2 ∣∣∣a=âLS=0 <strong>for</strong>i =1, 2,...,M. (31.91)∂a iClearly, if the measurement errors n i are indeed Gaussian distributed, as assumedabove, then the LS <strong>and</strong> ML estimators of the parameters a coincide. Becauseof its relative simplicity, the method of least squares is often applied to cases inwhich the n i are not Gaussian distributed. The resulting estimators â LS are not theML estimators, <strong>and</strong> the best that can be said in justification is that the method isan obviously sensible procedure <strong>for</strong> parameter estimation that has stood the testof time.Finally, we note that the method of least squares is easily extended to the casein which each measurement y i depends on several variables, which we denoteby x i . For example, y i might represent the temperature measured at the (threedimensional)position x i in a room. In this case, the data is modelled by afunction y = f(x i ; a), <strong>and</strong> the remainder of the above discussion carries throughunchanged.31.6.1 Linear least squaresWe have so far made no restriction on the <strong>for</strong>m of the function f(x; a). It sohappens, however, that, <strong>for</strong> a model in which f(x; a) isalinear function of theparameters a 1 ,a 2 ,...,a M , one can always obtain analytic expressions <strong>for</strong> the LSestimators â LS <strong>and</strong> their variances. The general <strong>for</strong>m of this kind of model isM∑f(x; a) = a i h i (x), (31.92)i=11272


31.6 THE METHOD OF LEAST SQUARESwhere {h 1 (x),h 2 (x),...,h M (x)} is some set of linearly independent fixed functionsof x, often called the basis functions. Note that the functions h i (x) themselves maybe highly non-linear functions of x. The ‘linear’ nature of the model (31.92) refersonly to its dependence on the parameters a i . Furthermore, in this case, it maybe shown that the LS estimators â i have zero bias <strong>and</strong> are minimum-variance,irrespective of the probability density function from which the measurement errorsn i are drawn.In order to obtain analytic expressions <strong>for</strong> the LS estimators â LS ,itisconvenientto write (31.92) in the <strong>for</strong>mf(x; a) =M∑R ij a j , (31.93)where R ij = h j (x i ) is an element of the response matrix R of the experiment. Theexpression <strong>for</strong> χ 2 given in (31.90) can then be written, in matrix notation, asj=1χ 2 (a) =(y − Ra) T N −1 (y − Ra). (31.94)The LS estimates of the parameters a are now found, as shown in (31.91), bydifferentiating (31.94) with respect to the a i <strong>and</strong> setting the resulting expressionsequal to zero. Denoting by ∇χ 2 the vector with elements ∂χ 2 /∂a i , we find∇χ 2 = −2R T N −1 (y − Ra). (31.95)This can be verified by writing out the expression (31.94) in component <strong>for</strong>m <strong>and</strong>differentiating directly.◮Verify result (31.95) by <strong>for</strong>mulating the calculation in component <strong>for</strong>m.To make the derivation less cumbersome, let us adopt the summation convention discussedin section 26.1, in which it is understood that any subscript that appears exactly twice inany term of an expression is to be summed over all the values that a subscript in thatposition can take. Thus, writing (31.94) in component <strong>for</strong>m, we haveχ 2 (a) =(y i − R ik a k )(N −1 ) ij (y j − R jl a l ).Differentiating with respect to a p gives∂χ 2= −R ik δ kp (N −1 ) ij (y j − R jl a l )+(y i − R ik a k )(N −1 ) ij (−R jl δ lp )∂a p= −R ip (N −1 ) ij (y j − R jl a l ) − (y i − R ik a k )(N −1 ) ij R jp , (31.96)where δ ij is the Kronecker delta symbol discussed in section 26.1. By swapping the indicesi <strong>and</strong> j in the second term on the RHS of (31.96) <strong>and</strong> using the fact that the matrix N −1is symmetric, we obtain∂χ 2= −2R ip (N −1 ) ij (y j − R jk a k )∂a p= −2(R T ) pi (N −1 ) ij (y j − R jk a k ). (31.97)If we denote the vector with components ∂χ 2 /∂a p , p =1, 2,...,M,by∇χ 2 <strong>and</strong> write theRHS of (31.97) in matrix notation, we recover the result (31.95). ◭1273


STATISTICSSetting the expression (31.95) equal to zero at a = â, we find−2R T N −1 y +2R T N −1 Râ =0.Provided the matrix R T N −1 R is not singular, we may solve this equation <strong>for</strong> â toobtainâ =(R T N −1 R) −1 R T N −1 y ≡ Sy, (31.98)thus defining the M×N matrix S. It follows that the LS estimates â i , i =1, 2,...,M,are linear functions of the original measurements y j , j =1, 2,...,N.Moreover,using the error propagation <strong>for</strong>mula (30.141) derived in subsection 30.12.3, wefind that the covariance matrix of the estimators â i is given byV ≡ Cov[â i , â j ]=SNS T =(R T N −1 R) −1 . (31.99)The two equations (31.98) <strong>and</strong> (31.99) contain the complete method of leastsquares. In particular, we note that, if one calculates the LS estimates using(31.98) then one has already obtained their covariance matrix (31.99).◮Prove result (31.99).Using the definition of S given in (31.98), the covariance matrix (31.99) becomesV = SNS T=[(R T N −1 R) −1 R T N −1 ]N[(R T N −1 R) −1 R T N −1 ] T .Using the result (AB ···C) T = C T ···B T A T <strong>for</strong> the transpose of a product of matrices <strong>and</strong>noting that, <strong>for</strong> any non-singular matrix, (A −1 ) T =(A T ) −1 we findV =(R T N −1 R) −1 R T N −1 N(N T ) −1 R[(R T N −1 R) T ] −1=(R T N −1 R) −1 R T N −1 R(R T N −1 R) −1=(R T N −1 R) −1 ,where we have also used the fact that N is symmetric <strong>and</strong> so N T = N. ◭It is worth noting that one may also write the elements of the (inverse)covariance matrix as(V −1 ) ij = 1 ( ∂ 2 χ 2,2 ∂a i ∂a j)a=âwhich is the same as the Fisher matrix (31.36) in cases where the measurementerrors are Gaussian distributed (<strong>and</strong> so the log-likelihood is ln L = −χ 2 /2). Thisproves,atleast<strong>for</strong>thiscase,ourearlierstatementthattheLSestimatorsareminimum-variance. In fact, since f(x; a) is linear in the parameters a, onecanwrite χ 2 exactly asχ 2 (a) =χ 2 (â)+ 1 M∑( ∂ 2 χ 2(a i − â i )(a j − â j ),2 ∂a i ∂a ji,j=1)a=âwhich is quadratic in the parameters a i . Hence the <strong>for</strong>m of the likelihood function1274


31.6 THE METHOD OF LEAST SQUARESy7654321001234 5xFigure 31.9 A set of data points with error bars indicating the uncertaintyσ =0.5 onthey-values. The straight line is y = ˆmx + ĉ, where ˆm <strong>and</strong> ĉ arethe least-squares estimates of the slope <strong>and</strong> intercept.L ∝ exp(−χ 2 /2) is Gaussian. From the discussions of subsections 31.3.6 <strong>and</strong>31.5.6, it follows that the ‘surfaces’ χ 2 (a) = c, where c is a constant, boundellipsoidal confidence regions <strong>for</strong> the parameters a i . The relationship between thevalue of the constant c <strong>and</strong> the confidence level is given by (31.39).◮An experiment produces the following data sample pairs (x i ,y i ):x i : 1.85 2.72 2.81 3.06 3.42 3.76 4.31 4.47 4.64 4.99y i : 2.26 3.10 3.80 4.11 4.74 4.31 5.24 4.03 5.69 6.57where the x i -values are known exactly but each y i -value is measured only to an accuracyof σ =0.5. Assuming the underlying model <strong>for</strong> the data to be a straight line y = mx + c,find the LS estimates of the slope m <strong>and</strong> intercept c <strong>and</strong> quote the st<strong>and</strong>ard error on eachestimate.The data are plotted in figure 31.9, together with error bars indicating the uncertainty inthe y i -values. Our model of the data is a straight line, <strong>and</strong> so we havef(x; c, m) =c + mx.In the language of (31.92), our basis functions are h 1 (x) =1<strong>and</strong>h 2 (x) =x <strong>and</strong> our modelparameters are a 1 = c <strong>and</strong> a 2 = m. From (31.93) the elements of the response matrix areR ij = h j (x i ), so that⎛ ⎞1 x 11 x 2R = ⎜⎝.⎟. ⎠ , (31.100)1 x Nwhere x i are the data values <strong>and</strong> N = 10 in our case. Further, since the st<strong>and</strong>ard deviationon each measurement error is σ, we have N = σ 2 I,whereI is the N × N identity matrix.Because of this simple <strong>for</strong>m <strong>for</strong> N, the expression (31.98) <strong>for</strong> the LS estimates reduces toâ = σ 2 (R T R) −1 1 σ 2 RT y =(R T R) −1 R T y. (31.101)Note that we cannot exp<strong>and</strong> the inverse in the last line, since R itself is not square <strong>and</strong>1275


STATISTICShence does not possess an inverse. Inserting the <strong>for</strong>m <strong>for</strong> R in (31.100) into the expression(31.101), we find( ) (ĉ=∑i 1 ∑i x ) −1 ( ∑∑iˆmi x ∑∑i y )ii i x2 ii x iy i( )( )1=x2−¯x Nȳ.N(x 2 − ¯x 2 ) −¯x 1 NxyWe thus obtain the LS estimatesxy − ¯x ȳˆm = <strong>and</strong> ĉ = x2 ȳ − ¯x xy= ȳ − ˆm¯x, (31.102)x 2 − ¯x 2 x 2 − ¯x 2where the last expression <strong>for</strong> ĉ shows that the best-fit line passes through the ‘centreof mass’ (¯x, ȳ) of the data sample. To find the st<strong>and</strong>ard errors on our results, we mustcalculate the covariance matrix of the estimators. This is given by (31.99), which in ourcase reduces toσ 2V = σ 2 (R T R) −1 =N(x 2 − ¯x 2 )( )x2−¯x. (31.103)−¯x 1The st<strong>and</strong>ard error on each estimator is simply the positive square root of the correspondingdiagonal element, i.e. σĉ = √ V 11 <strong>and</strong> σ ˆm = √ V 22 , <strong>and</strong> the covariance of the estimators ˆm<strong>and</strong> ĉ is given by Cov[ĉ, ˆm] =V 12 = V 21 . Inserting the data sample averages <strong>and</strong> momentsinto (31.102) <strong>and</strong> (31.103), we findc = ĉ ± σĉ =0.40 ± 0.62 <strong>and</strong> m = ˆm ± σ ˆm =1.11 ± 0.17.The ‘best-fit’ straight line y = ˆmx + ĉ is plotted in figure 31.9. For comparison, the truevalues used to create the data were m =1<strong>and</strong>c =1.◭The extension of the method to fitting data to a higher-order polynomial, suchas f(x; a) =a 1 + a 2 x + a 3 x 2 , is obvious. However, as the order of the polynomialincreases the matrix inversions become rather complicated. Indeed, even when thematrices are inverted numerically, the inversion is prone to numerical instabilities.A better approach is to replace the basis functions h m (x) =x m , m =1, 2,...,M,with a set of polynomials that are ‘orthogonal over the data’, i.e. such thatN∑h l (x i )h m (x i )=0 <strong>for</strong>l ≠ m.i=1Such a set of polynomial basis functions can always be found by using the Gram–Schmidt orthogonalisation procedure presented in section 17.1. The details of thisapproach are beyond the scope of our discussion but we note that, in this case,the matrix R T R is diagonal <strong>and</strong> may be inverted easily.31.6.2 Non-linear least squaresIf the function f(x; a) isnot linear in the parameters a then, in general, it isnot possible to obtain an explicit expression <strong>for</strong> the LS estimates â. Instead,onemust use an iterative (numerical) procedure, which we now outline. In practice,1276


31.7 HYPOTHESIS TESTINGhowever, such problems are best solved using one of the many commerciallyavailable software packages.One begins by making a first guess a 0 <strong>for</strong> the values of the parameters. At thispoint in parameter space, the components of the gradient ∇χ 2 will not be equalto zero, in general (unless one makes a very lucky guess!). Thus, <strong>for</strong> at least somevalues of i, we have∣∂χ 2 ∣∣∣a=a≠0.∂a i 0Our aim is to find a small increment δa in the values of the parameters, such that∣∂χ 2 ∣∣∣a=a0= 0 <strong>for</strong> all i. (31.104)∂a i +δaIf our first guess a 0 were sufficiently close to the true (local) minimum of χ 2 ,we could find the required increment δa by exp<strong>and</strong>ing the LHS of (31.104) as aTaylor series about a = a 0 , keeping only the zeroth-order <strong>and</strong> first-order terms:∣ ∣∂χ 2 ∣∣∣a=a0≈ ∂χ2 ∣∣∣a=a+∂a i +δa∂a i 0M∑j=1∣∂ 2 χ 2 ∣∣∣a=aδa j . (31.105)∂a i ∂a j 0Setting this expression to zero, we find that the increments δa j may be found bysolving the set of M linear equationsM∑j=1∣ ∣∂ 2 χ 2 ∣∣∣a=aδa j = − ∂χ2 ∣∣∣a=a.∂a i ∂a j 0 ∂a i 0It most cases, however, our first guess a 0 will not be sufficiently close to the trueminimum <strong>for</strong> (31.105) to be an accurate approximation, <strong>and</strong> consequently (31.104)will not be satisfied. In this case, a 1 = a 0 + δa is (hopefully) an improved guessat the parameter values; the whole process is then repeated until convergence isachieved.It is worth noting that, when one is estimating several parameters a, thefunction χ 2 (a) maybevery complicated. In particular, it may possess numerouslocal extrema. The procedure outlined above will converge to the local extremum‘nearest’ to the first guess a 0 . Since, in fact, we are interested only in the localminimum that has the absolute lowest value of χ 2 (a), it is clear that a large partof solving the problem is to make a ‘good’ first guess.31.7 Hypothesis testingSo far we have concentrated on using a data sample to obtain a number or a setof numbers. These numbers may be estimated values <strong>for</strong> the moments or centralmoments of the population from which the sample was drawn or, more generally,the values of some parameters a in an assumed model <strong>for</strong> the data. Sometimes,1277


STATISTICShowever, one wishes to use the data to give a ‘yes’ or ‘no’ answer to a particularquestion. For example, one might wish to know whether some assumed modeldoes, in fact, provide a good fit to the data, or whether two parameters have thesame value.31.7.1 Simple <strong>and</strong> composite hypothesesIn order to use data to answer questions of this sort, the question must beposed precisely. This is done by first asserting that some hypothesis is true.The hypothesis under consideration is traditionally called the null hypothesis<strong>and</strong> is denoted by H 0 . In particular, this usually specifies some <strong>for</strong>m P (x|H 0 )<strong>for</strong> the probability density function from which the data x are drawn. If thehypothesis determines the PDF uniquely, then it is said to be a simple hypothesis.If, however, the hypothesis determines the functional <strong>for</strong>m of the PDF but not thevalues of certain parameters a on which it depends then it is called a compositehypothesis.One decides whether to accept or reject the null hypothesis H 0 by per<strong>for</strong>mingsome statistical test, as described below in subsection 31.7.2. In fact, <strong>for</strong>mallyone uses a statistical test to decide between the null hypothesis H 0 <strong>and</strong> thealternative hypothesis H 1 . We define the latter to be the complement H 0 of thenull hypothesis within some restricted hypothesis space known (or assumed) inadvance. Hence, rejection of H 0 implies acceptance of H 1 , <strong>and</strong> vice versa.As an example, let us consider the case in which a sample x is drawn from aGaussian distribution with a known variance σ 2 but with an unknown mean µ.If one adopts the null hypothesis H 0 that µ = 0, which we write as H 0 : µ =0,then the corresponding alternative hypothesis must be H 1 : µ ≠ 0. Note that,in this case, H 0 is a simple hypothesis whereas H 1 is a composite hypothesis.If, however, one adopted the null hypothesis H 0 : µ


31.7 HYPOTHESIS TESTINGP (t|H 0 )αt crittP (t|H 1 )βt crittFigure 31.10 The sampling distributions P (t|H 0 )<strong>and</strong>P (t|H 1 ) of a test statistict. The shaded areas indicate the (one-tailed) regions <strong>for</strong> which Pr(t >t crit |H 0 )=α <strong>and</strong> Pr(t t crit |H 0 )= P (t|H 0 ) dt = α; (31.106)t critthis is indicated by the shaded region in the upper half of figure 31.10. Equally,a (one-tailed) rejection region could consist of values of t less than some valuet crit . Alternatively, one could define a (two-tailed) rejection region by two valuest 1 <strong>and</strong> t 2 such that Pr(t 1


STATISTICSfalse (in which case H 1 is true). The probability β (say)thatsuchanerrorwilloccur is, in general, difficult to calculate, since the alternative hypothesis H 1 isoften composite. Nevertheless, in the case where H 1 is a simple hypothesis, it isstraight<strong>for</strong>ward (in principle) to calculate β. Denoting the corresponding samplingdistribution of t by P (t|H 1 ), the probability β is the integral of P (t|H 1 )overthecomplement of the rejection region, called the acceptance region. For example, inthe case corresponding to (31.106) this probability is given byβ =Pr(tc, (31.107)P (t|H 1 )where c is some constant determined by the required significance level.In the case where the test statistic t is a simple scalar quantity, the Neyman–Pearson lemma is also useful in deciding which such statistic is the ‘best’ inthe sense of having the maximum power <strong>for</strong> a given significance level α. From(31.107), we can see that the best statistic is given by the likelihood ratiot(x) = P (x|H 0)P (x|H 1 ) . (31.108)<strong>and</strong> that the corresponding rejection region <strong>for</strong> H 0 is given by t


31.7 HYPOTHESIS TESTING◮Ten independent sample values x i , i =1, 2,...,10, are drawn at r<strong>and</strong>om from a Gaussi<strong>and</strong>istribution with st<strong>and</strong>ard deviation σ =1. The mean µ of the distribution is known toequal either zero or unity. The sample values are as follows:2.22 2.56 1.07 0.24 0.18 0.95 0.73 −0.79 2.09 1.81Test the null hypothesis H 0 : µ =0at the 10% significance level.The restricted nature of the hypothesis space means that our null <strong>and</strong> alternative hypothesesare H 0 : µ =0<strong>and</strong>H 1 : µ = 1 respectively. Since H 0 <strong>and</strong> H 1 are both simple hypotheses,the best test statistic is given by the likelihood ratio (31.108). Thus, denoting the meansby µ 0 <strong>and</strong> µ 1 , we havet(x) = exp [ − 1 2∑i (x i − µ 0 ) 2]exp [ ∑− 1 2 i (x i − µ 1 ) 2] = exp [ − 1 2∑i (x2 i − 2µ 0 x i + µ 2 0 )]exp [ − 1 2∑i (x2 i− 2µ 1 x i + µ 2 1 )]=exp [ (µ 0 − µ 1 ) ∑ i x i − 1 2 N(µ2 0 − µ 2 1) ] .Inserting the values µ 0 =0<strong>and</strong>µ 1 = 1, yields t =exp(−N¯x + 1 N), where ¯x is the2sample mean. Since − ln t is a monotonically decreasing function of t, however, we mayequivalently use as our test statisticv = − 1 N ln t + 1 2 = ¯x,where we have divided by the sample size N <strong>and</strong> added 1 <strong>for</strong> convenience. Thus we2may take the sample mean as our test statistic. From (31.13), we know that the samplingdistribution of the sample mean under our null hypothesis H 0 is the Gaussian distributionN(µ 0 ,σ 2 /N), where µ 0 =0,σ 2 =1<strong>and</strong>N = 10. Thus ¯x ∼ N(0, 0.1).Since ¯x is a monotonically decreasing function of t, our best rejection region <strong>for</strong> a givensignificance α is ¯x >¯x crit ,where¯x crit depends on α. Thus, in our case, ¯x crit is given by( )¯xcrit − µ 0α =1− Φ=1− Φ(10¯x crit ),σwhere Φ(z) is the cumulative distribution function <strong>for</strong> the st<strong>and</strong>ard Gaussian. For a 10%significance level we have α =0.1 <strong>and</strong>, from table 30.3 in subsection 30.9.1, we find¯x crit =0.128. Thus the rejection region on ¯x is¯x >0.128.From the sample, we deduce that ¯x =1.11, <strong>and</strong> so we can clearly reject the null hypothesisH 0 : µ = 0 at the 10% significance level. It can, in fact, be rejected at a much highersignificance level. As revealed on p. 1239, the data was generated using µ =1.◭31.7.4 The generalised likelihood-ratio testIf the null hypothesis H 0 or the alternative hypothesis H 1 is composite (or bothare composite) then the corresponding distributions P (x|H 0 )<strong>and</strong>P (x|H 1 )arenot uniquely determined, in general, <strong>and</strong> so we cannot use the Neyman–Pearsonlemma to obtain the ‘best’ test statistic t. Nevertheless, in many cases, there stillexists a general procedure <strong>for</strong> constructing a test statistic t which has useful1281


STATISTICSproperties <strong>and</strong> which reduces to the Neyman–Pearson statistic (31.108) in thespecial case where H 0 <strong>and</strong> H 1 are both simple hypotheses.Consider the quite general, <strong>and</strong> commonly occurring, case in which thedata sample x is drawn from a population P (x|a) with a known (or assumed)functional <strong>for</strong>m but depends on the unknown values of some parametersa 1 ,a 2 ,...,a M . Moreover, suppose we wish to test the null hypothesis H 0 thatthe parameter values a lie in some subspace S of the full parameter spaceA. In other words, on the basis of the sample x it is desired to test thenull hypothesis H 0 :(a 1 ,a 2 ,...,a M lies in S) against the alternative hypothesisH 1 :(a 1 ,a 2 ,...,a M lies in S), where S is A − S.Since the functional <strong>for</strong>m of the population is known, we may write down thelikelihood function L(x; a) <strong>for</strong> the sample. Ordinarily, the likelihood will havea maximum as the parameters a are varied over the entire parameter space A.This is the usual maximum-likelihood estimate of the parameter values, whichwe denote by â. If, however, the parameter values are allowed to vary only overthe subspace S then the likelihood function will be maximised at the point â S ,which may or may not coincide with the global maximum â. Now, let us take asour test statistic the generalised likelihood ratiot(x) = L(x; â S)L(x; â) , (31.109)where L(x; â S ) is the maximum value of the likelihood function in the subspaceS <strong>and</strong> L(x; â) is its maximum value in the entire parameter space A. Itisclearthat t is a function of the sample values only <strong>and</strong> must lie between 0 <strong>and</strong> 1.We will concentrate on the special case where H 0 is the simple hypothesisH 0 : a = a 0 . The subspace S then consists of only the single point a 0 . Thus(31.109) becomest(x) = L(x; a 0)L(x; â) , (31.110)<strong>and</strong> the sampling distribution P (t|H 0 ) can be determined (in principle). As in theprevious subsection, the best rejection region <strong>for</strong> a given significance α is simplyt


31.7 HYPOTHESIS TESTING◮Ten independent sample values x i , i =1, 2,...,10, are drawn at r<strong>and</strong>om from a Gaussi<strong>and</strong>istribution with st<strong>and</strong>ard deviation σ =1. The sample values are as follows:2.22 2.56 1.07 0.24 0.18 0.95 0.73 −0.79 2.09 1.81Test the null hypothesis H 0 : µ =0at the 10% significance level.We must test the (simple) null hypothesis H 0 : µ = 0 against the (composite) alternativehypothesis H 1 : µ ≠ 0. Thus, the subspace S is the single point µ =0,whereasA is theentire µ-axis. The likelihood function is1L(x; µ) =(2π) exp [ ∑− 1 N/2 2 i (x i − µ) 2] ,which has its global maximum at µ = ¯x. The test statistic t is then given byt(x) = L(x;0)L(x; ¯x) = exp [ ∑ ]− 1 2 i x2 iexp [ − 1 2∑i (x i − ¯x) 2] =exp( − 1 N¯x2) .2It is in fact more convenient to consider the test statisticv = −2lnt = N¯x 2 .Since −2lnt is a monotonically decreasing function of t, the rejection region now becomesv>v crit ,where∫ ∞P (v|H 0 ) dv = α, (31.111)v critα being the significance level of the test. Thus it only remains to determine the samplingdistribution P (v|H 0 ). Under the null hypothesis H 0 ,weexpect¯x to be Gaussian distributed,with mean zero <strong>and</strong> variance 1/N. Thus, from subsection 30.9.4, v will follow a chi-squareddistribution of order 1. Substituting the appropriate <strong>for</strong>m <strong>for</strong> P (v|H 0 ) in (31.111) <strong>and</strong> settingα =0.1, we find by numerical integration (or from table 31.2) that v crit = N¯x 2 crit =2.71.Since N = 10, the rejection region on ¯x at the 10% significance level is thus¯x 0.52.As noted be<strong>for</strong>e, <strong>for</strong> this sample ¯x =1.11, <strong>and</strong> so we may reject the null hypothesisH 0 : µ = 0 at the 10% significance level. ◭The above example illustrates the general situation that if the maximumlikelihoodestimates â of the parameters fall in or near the subspace S then thesample will be considered consistent with H 0 <strong>and</strong> the value of t will be nearunity. If â is distant from S then the sample will not be in accord with H 0 <strong>and</strong>ordinarily t will have a small (positive) value.It is clear that in order to prescribe the rejection region <strong>for</strong> t, or<strong>for</strong>arelatedstatistic u or v, it is necessary to know the sampling distribution P (t|H 0 ). If H 0is simple then one can in principle determine P (t|H 0 ), although this may provedifficult in practice. Moreover, if H 0 is composite, then it may not be possibleto obtain P (t|H 0 ), even in principle. Nevertheless, a useful approximate <strong>for</strong>m <strong>for</strong>P (t|H 0 ) exists in the large-sample limit. Consider the null hypothesis<strong>and</strong> the a 0 iH 0 :(a 1 = a 0 1,a 2 = a 0 2,...,a R = a 0 R),where R ≤ Mare fixed numbers. (In fact, we may fix the values of any subset1283


STATISTICScontaining R of the M parameters.) If H 0 is true then it follows from ourdiscussion in subsection 31.5.6 (although we shall not prove it) that, when thesample size N is large, the quantity −2lnt follows approximately a chi-squareddistribution of order R.31.7.5 Student’s t-testStudent’s t-test is just a special case of the generalised likelihood ratio test appliedto a sample x 1 ,x 2 ,...,x N drawn independently from a Gaussian distribution <strong>for</strong>which both the mean µ <strong>and</strong> variance σ 2 are unknown, <strong>and</strong> <strong>for</strong> which one wishesto distinguish between the hypothesesH 0 : µ = µ 0 , 0


31.7 HYPOTHESIS TESTINGThe sum of squares in the denominator of (31.112) may be put into the <strong>for</strong>m∑i (x i − µ 0 ) 2 = N(¯x − µ 0 ) 2 + ∑ i (x i − ¯x) 2 .Thus, on dividing the numerator <strong>and</strong> denominator in (31.112) by ∑ i (x i − ¯x) 2 <strong>and</strong>rearranging, the generalised likelihood ratio λ can be written) −N/2λ =(1+ t2,N − 1where we have defined the new variablet = ¯x − µ 0s/ √ N − 1 . (31.113)Since t 2 is a monotonically decreasing function of λ, the corresponding rejectionregion is t 2 > c, where c is a positive constant depending on the requiredsignificance level α. It is conventional, however, to use t itself as our test statistic,in which case our rejection region becomes two-tailed <strong>and</strong> is given bytt crit , (31.114)where t crit is the positive square root of the constant c.The definition (31.113) <strong>and</strong> the rejection region (31.114) <strong>for</strong>m the basis ofStudent’s t-test. It only remains to determine the sampling distribution P (t|H 0 ).At the outset, it is worth noting that if we write the expression (31.113) <strong>for</strong> tin terms of the st<strong>and</strong>ard estimator ˆσ = √ Ns 2 /(N − 1) of the st<strong>and</strong>ard deviationthen we obtaint = ¯x − µ 0ˆσ/ √ N . (31.115)If, in fact, we knew the true value of σ <strong>and</strong>useditinthisexpression<strong>for</strong>t thenit is clear from our discussion in section 31.3 that t would follow a Gaussi<strong>and</strong>istribution with mean 0 <strong>and</strong> variance 1, i.e. t ∼ N(0, 1). When σ is not known,however, we have to use our estimate ˆσ in (31.115), with the result that t isno longer distributed as the st<strong>and</strong>ard Gaussian. As one might expect fromthe central limit theorem, however, the distribution of t does tend towards thest<strong>and</strong>ard Gaussian <strong>for</strong> large values of N.As noted earlier, the exact distribution of t, valid <strong>for</strong> any value of N, wasfirstdiscovered by William Gossett. From (31.35), if the hypothesis H 0 is true then thejoint sampling distribution of ¯x <strong>and</strong> s is given by)P (¯x, s|H 0 )=Cs N−2 exp(− Ns22σ 2 exp[− N(¯x − µ 0) 2 ]2σ 2 ,(31.116)where C is a normalisation constant. We can use this result to obtain the jointsampling distribution of s <strong>and</strong> t by dem<strong>and</strong>ing thatP (¯x, s|H 0 ) d¯xds= P (t, s|H 0 ) dt ds.1285


STATISTICSUsing (31.113) to substitute <strong>for</strong> ¯x − µ 0 in (31.116), <strong>and</strong> noting that d¯x =(s/ √ N − 1) dt, we find( )]P (¯x, s|H 0 ) d¯xds= As N−1 exp[− Ns22σ 2 1+ t2dt ds,N − 1where A is another normalisation constant. In order to obtain the samplingdistribution of t alone, we must integrate P (t, s|H 0 ) with respect to s over itsallowed range, from 0 to ∞. Thus, the required distribution of t alone is given by∫ ∞∫ ∞( )]P (t|H 0 )= P (t, s|H 0 ) ds = A s N−1 exp[− Ns2002σ 2 1+ t2ds.N − 1(31.117)To carry out this integration, we set y = s{1+[t 2 /(N − 1)]} 1/2 , which on substitutioninto (31.117) yields) −N/2 ∫ ∞)P (t|H 0 )=A(1+ t2y N−1 exp(− Ny2N − 1 02σ 2 dy.Since the integral over y does not depend on t, it is simply a constant. We thusfind that that the sampling distribution of the variable t is1 Γ ( 12P (t|H 0 )= √ N) ((N − 1)π Γ ( 12 (N − 1))1+ t2N − 1) −N/2,(31.118)wherewehaveusedthecondition ∫ ∞−∞ P (t|H 0) dt = 1 to determine the normalisationconstant (see exercise 31.18).The distribution (31.118) is called Student’s t-distribution with N − 1 degrees offreedom. A plot of Student’s t-distribution is shown in figure 31.11 <strong>for</strong> variousvalues of N. For comparison, we also plot the st<strong>and</strong>ard Gaussian distribution,to which the t-distribution tends <strong>for</strong> large N. As is clear from the figure, thet-distribution is symmetric about t = 0. In table 31.3 we list some critical pointsof the cumulative probability function C n (t) ofthet-distribution, which is definedby∫ tC n (t) = P (t ′ |H 0 ) dt ′ ,−∞where n = N − 1 is the number of degrees of freedom. Clearly, C n (t) is analogousto the cumulative probability function Φ(z) of the Gaussian distribution, discussedin subsection 30.9.1. For comparison purposes, we also list the critical points ofΦ(z), which corresponds to the t-distribution <strong>for</strong> N = ∞.1286


31.7 HYPOTHESIS TESTINGP (t|H 0 )0.50.40.30.2N =3N =2N =10N =50.10−4 −3 −2 −1 0 1 2 3 4tFigure 31.11 Student’s t-distribution <strong>for</strong> various values of N. The brokencurve shows the st<strong>and</strong>ard Gaussian distribution <strong>for</strong> comparison.◮Ten independent sample values x i , i =1, 2,...,10, are drawn at r<strong>and</strong>om from a Gaussi<strong>and</strong>istribution with unknown mean µ <strong>and</strong> unknown st<strong>and</strong>ard deviation σ. The sample valuesare as follows:2.22 2.56 1.07 0.24 0.18 0.95 0.73 −0.79 2.09 1.81Test the null hypothesis H 0 : µ =0at the 10% significance level.For our null hypothesis, µ 0 = 0. Since <strong>for</strong> this sample ¯x =1.11, s =1.01 <strong>and</strong> N = 10, itfollows from (31.113) that¯xt =s/ √ N − 1 =3.33.The rejection region <strong>for</strong> t is given by (31.114) where t crit is such thatC N−1 (t crit )=1− α/2,<strong>and</strong> α is the required significance of the test. In our case α =0.1 <strong>and</strong>N = 10, <strong>and</strong> fromtable 31.3 we find t crit =1.83. Thus our rejection region <strong>for</strong> H 0 at the 10% significancelevel ist1.83.For our sample t =3.30 <strong>and</strong> so we can clearly reject the null hypothesis H 0 : µ =0atthislevel. ◭It is worth noting the connection between the t-test <strong>and</strong> the classical confidenceinterval on the mean µ. The central confidence interval on µ at the confidencelevel 1 − α is the set of values <strong>for</strong> which−t crit < ¯x − µs/ √ N − 1


STATISTICSC n (t) 0.5 0.6 0.7 0.8 0.9 0.950 0.975 0.990 0.995 0.999n =1 0.00 0.33 0.73 1.38 3.08 6.31 12.7 31.8 63.7 318.32 0.00 0.29 0.62 1.06 1.89 2.92 4.30 6.97 9.93 22.33 0.00 0.28 0.58 0.98 1.64 2.35 3.18 4.54 5.84 10.24 0.00 0.27 0.57 0.94 1.53 2.13 2.78 3.75 4.60 7.175 0.00 0.27 0.56 0.92 1.48 2.02 2.57 3.37 4.03 5.896 0.00 0.27 0.55 0.91 1.44 1.94 2.45 3.14 3.71 5.217 0.00 0.26 0.55 0.90 1.42 1.90 2.37 3.00 3.50 4.798 0.00 0.26 0.55 0.89 1.40 1.86 2.31 2.90 3.36 4.509 0.00 0.26 0.54 0.88 1.38 1.83 2.26 2.82 3.25 4.3010 0.00 0.26 0.54 0.88 1.37 1.81 2.23 2.76 3.17 4.1411 0.00 0.26 0.54 0.88 1.36 1.80 2.20 2.72 3.11 4.0312 0.00 0.26 0.54 0.87 1.36 1.78 2.18 2.68 3.06 3.9313 0.00 0.26 0.54 0.87 1.35 1.77 2.16 2.65 3.01 3.8514 0.00 0.26 0.54 0.87 1.35 1.76 2.15 2.62 2.98 3.7915 0.00 0.26 0.54 0.87 1.34 1.75 2.13 2.60 2.95 3.7316 0.00 0.26 0.54 0.87 1.34 1.75 2.12 2.58 2.92 3.6917 0.00 0.26 0.53 0.86 1.33 1.74 2.11 2.57 2.90 3.6518 0.00 0.26 0.53 0.86 1.33 1.73 2.10 2.55 2.88 3.6119 0.00 0.26 0.53 0.86 1.33 1.73 2.09 2.54 2.86 3.5820 0.00 0.26 0.53 0.86 1.33 1.73 2.09 2.53 2.85 3.5525 0.00 0.26 0.53 0.86 1.32 1.71 2.06 2.49 2.79 3.4630 0.00 0.26 0.53 0.85 1.31 1.70 2.04 2.46 2.75 3.3940 0.00 0.26 0.53 0.85 1.30 1.68 2.02 2.42 2.70 3.3150 0.00 0.26 0.53 0.85 1.30 1.68 2.01 2.40 2.68 3.26100 0.00 0.25 0.53 0.85 1.29 1.66 1.98 2.37 2.63 3.17200 0.00 0.25 0.53 0.84 1.29 1.65 1.97 2.35 2.60 3.13∞ 0.00 0.25 0.52 0.84 1.28 1.65 1.96 2.33 2.58 3.09Table 31.3 The confidence limits t of the cumulative probability functionC n (t) <strong>for</strong> Student’s t-distribution with n degrees of freedom. For example,C 5 (0.92) = 0.8. The row n = ∞ is also the corresponding result <strong>for</strong> thest<strong>and</strong>ard Gaussian distribution.where t crit satisfies C N−1 (t crit )=α/2. Thus the required confidence interval is¯x −√ t crits< µ < ¯x + √ t crits.N − 1 N − 1Hence, in the above example, the 90% classical central confidence interval on µis0.49


31.7 HYPOTHESIS TESTINGdistributions. In particular, let us consider the case where we have two independentsamples of sizes N 1 <strong>and</strong> N 2 , drawn respectively from Gaussian distributions witha common variance σ 2 but with possibly different means µ 1 <strong>and</strong> µ 2 . On the basisof the samples, one wishes to distinguish between the hypothesesH 0 : µ 1 = µ 2 , 0


STATISTICSwhere α is the required significance level of the test. In our case we set α =0.05, <strong>and</strong> fromtable 31.3 with n = 16 we find that t crit =2.12. The rejection region is there<strong>for</strong>et2.12.Since t = −2.17 <strong>for</strong> our samples, we can reject the null hypothesis H 0 : µ 1 = µ 2 , althoughonly by a small margin. (Indeed, it is easily shown that one cannot reject H 0 at the 2%significance level). The 95% central confidence interval on ω = µ 1 − µ 2 is given by( ) 1/2 ( ) 1/2 N1 + N 2N1 + N 2¯w − ˆσt crit < ω < ¯w + ˆσt crit ,N 1 N 2 N 1 N 2where t crit is given by (31.120). Thus, we find−26.1


31.7 HYPOTHESIS TESTINGλ(u)0.100.0500λ crita b10 20 30 40uFigure 31.12 The sampling distribution P (u|H 0 )<strong>for</strong>N = 10; this is a chisquareddistribution <strong>for</strong> N − 1 degrees of freedom.distribution with unknown µ <strong>and</strong> σ, <strong>and</strong> we wish to distinguish between the twohypothesesH 0 : σ 2 = σ 2 0, −∞


STATISTICShowever, it is difficult to determine a <strong>and</strong> b <strong>for</strong> a given significance level α, soaslightly different rejection region, which we now describe, is usually adopted.The sampling distribution P (u|H 0 ) may be found straight<strong>for</strong>wardly from thesampling distribution of s given in (31.35). Let us first determine P (s 2 |H 0 )bydem<strong>and</strong>ing thatfrom which we findP (s 2 |H 0 )= P (s|H 0)2sP (s|H 0 ) ds = P (s 2 |H 0 ) d(s 2 ),( ) (N−1)/2 N (s 2 ) (N−3)/2 ( )=2σ02 Γ ( 12 (N − 1)) exp − Ns22σ02 . (31.122)Thus, the sampling distribution of u = Ns 2 /σ0 2 is given byP (u|H 0 )=12 (N−1)/2 Γ ( 12 (N − 1))u(N−3)/2 exp ( − 1 2 u) .We note, in passing, that the distribution of u is precisely that of an (N − 1) thorderchi-squared variable (see subsection 30.9.4), i.e. u ∼ χ 2 N−1 . Although it doesnot give quite the best test, one then takes the rejection region to be0


31.7 HYPOTHESIS TESTINGWe now turn to Fisher’s F-test. Let us suppose that two independent samplesof sizes N 1 <strong>and</strong> N 2 are drawn from Gaussian distributions with means <strong>and</strong>variances µ 1 ,σ1 2 <strong>and</strong> µ 2,σ2 2 respectively, <strong>and</strong> we wish to distinguish between thetwo hypothesesH 0 : σ 2 1 = σ 2 2 <strong>and</strong> H 1 : σ 2 1 ≠ σ 2 2.In this case, the generalised likelihood ratio is found to beλ = (N [1 + N 2 ) (N1+N2)/2 F(N1 − 1)/(N 2 − 1) ] N 1/2N N1/21N N2/2 [2 1+F(N1 − 1)/(N 2 − 1) ] ,(N 1+N 2)/2where F is given by the variance ratioF = N 1s 2 1 /(N 1 − 1)N 2 s 2 2 /(N 2 − 1) ≡ u2v 2 (31.123)<strong>and</strong> s 1 <strong>and</strong> s 2 are the st<strong>and</strong>ard deviations of the two samples. On plotting λ as afunction of F, it is apparent that the rejection region λ


STATISTICSdistribution P (v 2 ,F|H 0 ) is obtained by requiring thatP (v 2 ,F|H 0 ) d(v 2 ) dF = P (u 2 |H 0 )P (v 2 |H 0 ) d(u 2 ) d(v 2 ).(31.125)In order to find the distribution of F alone, we now integrate P (v 2 ,F|H 0 ) withrespect to v 2 from 0 to ∞, from which we obtainP (F|H 0 )(N1 − 1=N 2 − 1) (N1−1)/2B ( 1F (N1−3)/2 (2 (N 1 − 1), 1 2 (N 2 − 1) )1+ N ) −(N1+N1 − 12−2)/2N 2 − 1 F ,(31.126)where B ( 12 (N 1 − 1), 1 2 (N 2 − 1) ) is the beta function defined in the Appendix.P (F|H 0 ) is called the F-distribution (or occasionally the Fisher distribution) with(N 1 − 1, N 2 − 1) degrees of freedom.◮Evaluate the integral ∫ ∞0 P (v2 ,F|H 0 ) d(v 2 ) to obtain result (31.126).From (31.125), we haveP (F|H 0 )=AF (N 1−3)/2∫ ∞0(v 2 ) (N 1+N 2 −4)/2 exp{− [(N }1 − 1)F +(N 2 − 1)]v 2d(v 2 ).2σ 2Making the substitution x =[(N 1 − 1)F +(N 2 − 1)]v 2 /(2σ 2 ), we obtain[P (F|H 0 )=A2σ 2(N 1 − 1)F +(N 2 − 1)[2σ 2= A(N 1 − 1)F +(N 2 − 1)] (N1 +N 2 −2)/2F (N 1−3)/2∫ ∞0x (N 1+N 2 −4)/2 e −x dx] (N1 +N 2 −2)/2F (N 1−3)/2 Γ ( 12 (N 1 + N 2 − 2) ) ,where in the last line we have used the definition of the gamma function given in theAppendix. Using the further result (18.165), which expresses the beta function in terms ofthe gamma function, <strong>and</strong> the expression <strong>for</strong> A given in (31.124), we see that P (F|H 0 )isindeed given by (31.126). ◭As it does not matter whether the ratio F given in (31.123) is defined as u 2 /v 2or as v 2 /u 2 , it is conventional to put the larger sample variance on the top, sothat F is always greater than or equal to unity. A large value of F indicates thatthe sample variances u 2 <strong>and</strong> v 2 are very different whereas a value of F close tounity means that they are very similar. There<strong>for</strong>e, <strong>for</strong> a given significance α, itis1294


31.7 HYPOTHESIS TESTINGC n1 ,n 2(F) n 1 = 1 2 3 4 5 6 7 8n 2 =1 161 200 216 225 230 234 237 2392 18.5 19.0 19.2 19.2 19.3 19.3 19.4 19.43 10.1 9.55 9.28 9.12 9.01 8.94 8.89 8.854 7.71 6.94 6.59 6.39 6.26 6.16 6.09 6.045 6.61 5.79 5.41 5.19 5.05 4.95 4.88 4.826 5.99 5.14 4.76 4.53 4.39 4.28 4.21 4.157 5.59 4.74 4.35 4.12 3.97 3.87 3.79 3.738 5.32 4.46 4.07 3.84 3.69 3.58 3.50 3.449 5.12 4.26 3.86 3.63 3.48 3.37 3.29 3.2310 4.96 4.10 3.71 3.48 3.33 3.22 3.14 3.0720 4.35 3.49 3.10 2.87 2.71 2.60 2.51 2.4530 4.17 3.32 2.92 2.69 2.53 2.42 2.33 2.2740 4.08 3.23 2.84 2.61 2.45 2.34 2.25 2.1850 4.03 3.18 2.79 2.56 2.40 2.29 2.20 2.13100 3.94 3.09 2.70 2.46 2.31 2.19 2.10 2.03∞ 3.84 3.00 2.60 2.37 2.21 2.10 2.01 1.94n 1 = 9 10 20 30 40 50 100 ∞n 2 =1 241 242 248 250 251 252 253 2542 19.4 19.4 19.4 19.5 19.5 19.5 19.5 19.53 8.81 8.79 8.66 8.62 8.59 8.58 8.55 8.534 6.00 5.96 5.80 5.75 5.72 5.70 5.66 5.635 4.77 4.74 4.56 4.50 4.46 4.44 4.41 4.376 4.10 4.06 3.87 3.81 3.77 3.75 3.71 3.677 3.68 3.64 3.44 3.38 3.34 3.32 3.27 3.238 3.39 3.35 3.15 3.08 3.04 3.02 2.97 2.939 3.18 3.14 2.94 2.86 2.83 2.80 2.76 2.7110 3.02 2.98 2.77 2.70 2.66 2.64 2.59 2.5420 2.39 2.35 2.12 2.04 1.99 1.97 1.91 1.8430 2.21 2.16 1.93 2.69 1.79 1.76 1.70 1.6240 2.12 2.08 1.84 1.74 1.69 1.66 1.59 1.5150 2.07 2.03 1.78 1.69 1.63 1.60 1.52 1.44100 1.97 1.93 1.68 1.57 1.52 1.48 1.39 1.28∞ 1.88 1.83 1.57 1.46 1.39 1.35 1.24 1.00Table 31.4 Values of F <strong>for</strong> which the cumulative probability function C n1 ,n 2(F)of the F-distribution with (n 1 ,n 2 ) degrees of freedom has the value 0.95. Forexample, <strong>for</strong> n 1 =10<strong>and</strong>n 2 =6,C n1 ,n 2(4.06) = 0.95.customary to define the rejection region on F as F>F crit ,whereC n1,n 2(F crit )=∫ Fcrit1P (F|H 0 ) dF = α,<strong>and</strong> n 1 = N 1 − 1 <strong>and</strong> n 2 = N 2 − 1 are the numbers of degrees of freedom.Table 31.4 lists values of F crit corresponding to the 5% significance level (i.e.α =0.05) <strong>for</strong> various values of n 1 <strong>and</strong> n 2 .1295


STATISTICS◮Suppose that two classes of students take the same mathematics examination <strong>and</strong> thefollowing percentage marks are obtained:Class 1: 66 62 34 55 77 80 55 60 69 47 50Class 2: 64 90 76 56 81 72 70Assuming that the two sets of examinations marks are drawn from Gaussian distributions,test the hypothesis H 0 : σ1 2 = σ2 2at the 5% significance level.The variances of the two samples are s 2 1 =(12.8)2 <strong>and</strong> s 2 2 =(10.3)2 <strong>and</strong> the sample sizesare N 1 =11<strong>and</strong>N 2 = 7. Thus, we haveu 2 = N 1s 2 1N 1 − 1 = 180.2 <strong>and</strong> v2 = N 2s 2 2N 2 − 1 = 123.8,where we have taken u 2 to be the larger value. Thus, F = u 2 /v 2 =1.46 to two decimalplaces. Since the first sample contains eleven values <strong>and</strong> the second contains seven values,we take n 1 =10<strong>and</strong>n 2 = 6. Consulting table 31.4, we see that, at the 5% significancelevel, F crit =4.06. Since our value lies com<strong>for</strong>tably below this, we conclude that there isno statistical evidence <strong>for</strong> rejecting the hypothesis that the two samples were drawn fromGaussian distributions with a common variance. ◭It is also common to define the variable z = 1 2ln F, the distribution of whichcan be found straightfowardly from (31.126). This is a useful change of variablesince it can be shown that, <strong>for</strong> large values of n 1 <strong>and</strong> n 2 , the variable z isdistributed approximately as a Gaussian with mean 1 2 (n−1 2 − n −11 ) <strong>and</strong> variance12 (n−1 2 + n −11 ). 31.7.7 Goodness of fit in least-squares problemsWe conclude our discussion of hypothesis testing with an example of a goodnessof-fittest. In section 31.6, we discussed the use of the method of least squares inestimating the best-fit values of a set of parameters a in a given model y = f(x; a)<strong>for</strong> a data set (x i ,y i ), i =1, 2,...,N. We have not addressed, however, the questionof whether the best-fit model y = f(x; â) does, in fact, provide a good fit to thedata. In other words, we have not considered thus far how to verify that thefunctional <strong>for</strong>m f of our assumed model is indeed correct. In the language ofhypothesis testing, we wish to distinguish between the two hypothesesH 0 : model is correct <strong>and</strong> H 1 : model is incorrect.Given the vague nature of the alternative hypothesis H 1 , we clearly cannot usethe generalised likelihood-ratio test. Nevertheless, it is still possible to test thenull hypothesis H 0 at a given significance level α.The least-squares estimates of the parameters â 1 , â 2 ,...,â M , as discussed insection 31.6, are those values that minimise the quantityN∑χ 2 (a) = [y i − f(x i ; a)](N −1 ) ij [y j − f(x j ; a)] = (y − f) T N −1 (y − f).i,j=11296


31.7 HYPOTHESIS TESTINGIn the last equality, we rewrote the expression in matrix notation by defining thecolumn vector f with elements f i = f(x i ; a). The value χ 2 (â) at this minimum canbe used as a statistic to test the null hypothesis H 0 , as follows. The N quantitiesy i − f(x i ; a) are Gaussian distributed. However, provided the function f(x j ; a) islinear in the parameters a, the equations (31.98) that determine the least-squaresestimate â constitute a set of M linear constraints on these N quantities. Thus,as discussed in subsection 30.15.2, the sampling distribution of the quantity χ 2 (â)will be a chi-squared distribution with N − M degrees of freedom (d.o.f), which hasthe expectation value <strong>and</strong> varianceE[χ 2 (â)] = N − M <strong>and</strong> V [χ 2 (â)] = 2(N − M).Thus we would expect the value of χ 2 (â) to lie typically in the range (N − M) ±√ 2(N − M). A value lying outside this range may suggest that the assumed model<strong>for</strong> the data is incorrect. A very small value of χ 2 (â) is usually an indication thatthe model has too many free parameters <strong>and</strong> has ‘over-fitted’ the data. Morecommonly, the assumed model is simply incorrect, <strong>and</strong> this usually results in avalue of χ 2 (â) that is larger than expected.One can choose to per<strong>for</strong>m either a one-tailed or a two-tailed test on thevalue of χ 2 (â). It is usual, <strong>for</strong> a given significance level α, to define the one-tailedrejection region to be χ 2 (â) >k,wheretheconstantk satisfies∫ ∞kP (χ 2 n ) dχ2 n = α (31.127)<strong>and</strong> P (χ 2 n) is the PDF of the chi-squared distribution with n = N − M degrees offreedom (see subsection 30.9.4).◮An experiment produces the following data sample pairs (x i ,y i ):x i : 1.85 2.72 2.81 3.06 3.42 3.76 4.31 4.47 4.64 4.99y i : 2.26 3.10 3.80 4.11 4.74 4.31 5.24 4.03 5.69 6.57where the x i -values are known exactly but each y i -value is measured only to an accuracyof σ =0.5. At the one-tailed 5% significance level, test the null hypothesis H 0 that theunderlying model <strong>for</strong> the data is a straight line y = mx + c.These data are the same as those investigated in section 31.6 <strong>and</strong> plotted in figure 31.9. Asshown previously, the least squares estimates of the slope m <strong>and</strong> intercept c are given byˆm =1.11 <strong>and</strong> ĉ =0.4. (31.128)Since the error on each y i -value is drawn independently from a Gaussian distribution withst<strong>and</strong>ard deviation σ, we haveN∑[ ] 2χ 2 yi − f(x i ; a)N∑ [ yi − mx i − c] 2(a) ==. (31.129)σσi=1i=1Inserting the values (31.128) into (31.129), we obtain χ 2 ( ˆm, ĉ) =11.5. In our case, thenumber of data points is N = 10 <strong>and</strong> the number of fitted parameters is M = 2. Thus, the1297


STATISTICSnumber of degrees of freedom is n = N − M = 8. Setting n =8<strong>and</strong>α =0.05 in (31.127)we find from table 31.2 that k =15.51. Hence our rejection region isχ 2 ( ˆm, ĉ) > 15.51.Since above we found χ 2 ( ˆm, ĉ) = 11.5, we cannot reject the null hypothesis that theunderlying model <strong>for</strong> the data is a straight line y = mx + c. ◭As mentioned above, our analysis is only valid if the function f(x; a) is linearin the parameters a. Nevertheless, it is so convenient that it is sometimes appliedin non-linear cases, provided the non-linearity is not too severe.31.8 Exercises31.1 A group of students uses a pendulum experiment to measure g, the accelerationof free fall, <strong>and</strong> obtains the following values (in m s −2 ): 9.80, 9.84, 9.72, 9.74,9.87, 9.77, 9.28, 9.86, 9.81, 9.79, 9.82. What would you give as the best value <strong>and</strong>st<strong>and</strong>ard error <strong>for</strong> g as measured by the group?31.2 Measurements of a certain quantity gave the following values: 296, 316, 307, 278,312, 317, 314, 307, 313, 306, 320, 309. Within what limits would you say there isa 50% chance that the correct value lies?31.3 The following are the values obtained by a class of 14 students when measuringa physical quantity x: 53.8, 53.1, 56.9, 54.7, 58.2, 54.1, 56.4, 54.8, 57.3, 51.0, 55.1,55.0, 54.2, 56.6.(a) Display these results as a histogram <strong>and</strong> state what you would give as thebest value <strong>for</strong> x.(b) Without calculation, estimate how much reliance could be placed upon youranswer to (a).(c) Databooks give the value of x as 53.6 with negligible error. Are the dataobtained by the students in conflict with this?31.4 Two physical quantities x <strong>and</strong> y are connected by the equationy 1/2 x=ax 1/2 + b ,<strong>and</strong> measured pairs of values <strong>for</strong> x <strong>and</strong> y are as follows:x: 10 12 16 20y: 409 196 114 94Determine the best values <strong>for</strong> a <strong>and</strong> b by graphical means, <strong>and</strong> (either by h<strong>and</strong>or by using a built-in calculator routine) by a least-squares fit to an appropriatestraight line.31.5 Measured quantities x <strong>and</strong> y are known to be connected by the <strong>for</strong>mulay =axx 2 + b ,where a <strong>and</strong> b are constants. Pairs of values obtained experimentally arex: 2.0 3.0 4.0 5.0 6.0y: 0.32 0.29 0.25 0.21 0.18Use these data to make best estimates of the values of y that would be obtained<strong>for</strong> (a) x =7.0, <strong>and</strong> (b) x = −3.5. As measured by fractional error, which estimateis likely to be the more accurate?1298


31.8 EXERCISES31.6 Prove that the sample mean is the best linear unbiased estimator of the populationmean µ as follows.(a) If the real numbers a 1 ,a 2 ,...,a n satisfy the constraint ∑ ni=1 a i = C, whereCis a given constant, show that ∑ ni=1 a2 i is minimised by a i = C/n <strong>for</strong> all i.(b) Consider the linear estimator ˆµ = ∑ ni=1 a ix i . Impose the conditions (i) that itis unbiased <strong>and</strong> (ii) that it is as efficient as possible.31.7 A population contains individuals of k types in equal proportions. A quantity Xhas mean µ i amongst individuals of type i <strong>and</strong> variance σ 2 , which has the samevalue <strong>for</strong> all types. In order to estimate the mean of X over the whole population,two schemes are considered; each involves a total sample size of nk. In the firstthe sample is drawn r<strong>and</strong>omly from the whole population, whilst in the second(stratified sampling) n individuals are r<strong>and</strong>omly selected from each of the k types.Show that in both cases the estimate has expectationµ = 1 kk∑µ i ,i=1but that the variance of the first scheme exceeds that of the second by an amount1k∑(µk 2 i − µ) 2 .n31.8 Carry through the following proofs of statements made in subsections 31.5.2 <strong>and</strong>31.5.3 about the ML estimators ˆτ <strong>and</strong> ˆλ.i=1(a) Find the expectation values of the ML estimators ˆτ <strong>and</strong> ˆλ given, respectively,in (31.71) <strong>and</strong> (31.75). Hence verify equations (31.76), which show that, eventhough an ML estimator is unbiased, it does not follow that functions of itare also unbiased.(b) Show that E[ˆτ 2 ]=(N+1)τ 2 /N <strong>and</strong> hence prove that ˆτ is a minimum-varianceestimator of τ.31.9 Each of a series of experiments consists of a large, but unknown, number n(≫ 1) of trials in each of which the probability of success p is the same, but alsounknown. In the ith experiment, i =1, 2,...,N, the total number of successes isx i (≫ 1). Determine the log-likelihood function.Using Stirling’s approximation to ln(n − x), show thatd ln(n − x)dn<strong>and</strong> hence evaluate ∂( n C x )/∂n.≈1+ln(n − x),2(n − x)By finding the (coupled) equations determining the ML estimators ˆp <strong>and</strong> ˆn,show that, to order n −1 , they must satisfy the simultaneous ‘arithmetic’ <strong>and</strong>‘geometric’ mean constraintsˆnˆp = 1 NN∑N∏ (x i <strong>and</strong> (1 − ˆp) N =i=1i=11 − x iˆn).1299


STATISTICS31.10 This exercise is intended to illustrate the dangers of applying <strong>for</strong>malised estimatortechniques to distributions that are not well behaved in a statistical sense.The following are five sets of 10 values, all drawn from the same Cauchydistribution with parameter a.(i) 4.81 −1.24 1.30 −0.23 2.98−1.13 −8.32 2.62 −0.79 −2.85(ii) 0.07 1.54 0.38 −2.76 −8.821.86 −4.75 4.81 1.14 −0.66(iii) 0.72 4.57 0.86 −3.86 0.30−2.00 2.65 −17.44 −2.26 −8.83(iv) −0.15 202.76 −0.21 −0.58 −0.140.36 0.44 3.36 −2.96 5.51(v) 0.24 −3.33 −1.30 3.05 3.991.59 −7.76 0.91 2.80 −6.46Ignoring the fact that the Cauchy distribution does not have a finite variance (oreven a <strong>for</strong>mal mean), show that â, theMLestimatorofa, has to satisfy∑101s(â) ==5.1+x 2 i=1 i/â2(∗)Using a programmable calculator, spreadsheet or computer, find the value ofâ that satisfies (∗) <strong>for</strong> each of the data sets <strong>and</strong> compare it with the valuea =1.6 used to generate the data. Form an opinion regarding the variance of theestimator.Show further that if it is assumed that (E[â]) 2 = E[â 2 ], then E[â] =ν 1/22,whereν 2 is the second (central) moment of the distribution, which <strong>for</strong> the Cauchydistribution is infinite!31.11 According to a particular theory, two dimensionless quantities X <strong>and</strong> Y haveequal values. Nine measurements of X gave values of 22, 11, 19, 19, 14, 27, 8,24 <strong>and</strong> 18, whilst seven measured values of Y were 11, 14, 17, 14, 19, 16 <strong>and</strong>14. Assuming that the measurements of both quantities are Gaussian distributedwith a common variance, are they consistent with the theory? An alternativetheory predicts that Y 2 = π 2 X; are the data consistent with this proposal?31.12 On a certain (testing) steeplechase course there are 12 fences to be jumped, <strong>and</strong>any horse that falls is not allowed to continue in the race. In a season of racinga total of 500 horses started the course <strong>and</strong> the following numbers fell at eachfence:Fence: 1 2 3 4 5 6 7 8 9 10 11 12Falls: 62 75 49 29 33 25 30 17 19 11 15 12Use this data to determine the overall probability of a horse’s falling at a fence,<strong>and</strong> test the hypothesis that it is the same <strong>for</strong> all horses <strong>and</strong> fences as follows.(a) Draw up a table of the expected number of falls at each fence on the basisof the hypothesis.(b) Consider <strong>for</strong> each fence i the st<strong>and</strong>ardised variableestimated falls − actual fallsz i =st<strong>and</strong>ard deviation of estimated falls ,<strong>and</strong> use it in an appropriate χ 2 test.(c) Show that the data indicates that the odds against all fences being equallytesting are about 40 to 1. Identify the fences that are significantly easier orharder than the average.1300


31.8 EXERCISES31.13 A similar technique to that employed in exercise 31.12 can be used to testcorrelations between characteristics of sampled data. To illustrate this considerthe following problem.During an investigation into possible links between mathematics <strong>and</strong> classicalmusic, pupils at a school were asked whether they had preferences (a) betweenmathematics <strong>and</strong> english, <strong>and</strong> (b) between classical <strong>and</strong> pop music. The resultsare given below.Classical None PopMathematics 23 13 14None 17 17 36English 30 10 40By computing tables of expected numbers, based on the assumption that nocorrelations exist, <strong>and</strong> calculating the relevant values of χ 2 , determine whetherthere is any evidence <strong>for</strong>(a) a link between academic <strong>and</strong> musical tastes, <strong>and</strong>(b) a claim that pupils either had preferences in both areas or had no preference.You will need to consider the appropriate value <strong>for</strong> the number of degrees offreedom to use when applying the χ 2 test.31.14 Three c<strong>and</strong>idates X, Y <strong>and</strong> Z were st<strong>and</strong>ing <strong>for</strong> election to a vacant seat ontheir college’s Student Committee. The members of the electorate (current firstyearstudents, consisting of 150 men <strong>and</strong> 105 women) were each allowed tocross out the name of the c<strong>and</strong>idate they least wished to be elected, the othertwo c<strong>and</strong>idates then being credited with one vote each. The following data areknown.(a) X received 100 votes from men, whilst Y received 65 votes from women.(b) Z received five more votes from men than X received from women.(c) The total votes cast <strong>for</strong> X <strong>and</strong> Y were equal.Analyse this data in such a way that a χ 2 test can be used to determine whethervoting was other than r<strong>and</strong>om (i) amongst men <strong>and</strong> (ii) amongst women.31.15 A particle detector consisting of a shielded scintillator is being tested by placing itnear a particle source whose intensity can be controlled by the use of absorbers.It might register counts even in the absence of particles from the source becauseof the cosmic ray background.The number of counts n registered in a fixed time interval as a function of thesource strength s is given in as:source strength s: 0 1 2 3 4 5 6counts n: 6 11 20 42 44 62 61At any given source strength, the number of counts is expected to be Poissondistributed with meann = a + bs,where a <strong>and</strong> b are constants. Analyse the data <strong>for</strong> a fit to this relationship <strong>and</strong>obtain the best values <strong>for</strong> a <strong>and</strong> b together with their st<strong>and</strong>ard errors.(a) How well is the cosmic ray background determined?(b) What is the value of the correlation coefficient between a <strong>and</strong> b? Isthisconsistent with what would happen if the cosmic ray background wereimagined to be negligible?(c) Do the data fit the expected relationship well? Is there any evidence that thereported data ‘are too good a fit’?1301


STATISTICS31.16 The function y(x) is known to be a quadratic function of x. The following tablegives the measured values <strong>and</strong> uncorrelated st<strong>and</strong>ard errors of y measured atvarious values of x (in which there is negligible error):x 1 2 3 4 5y(x) 3.5 ± 0.5 2.0 ± 0.5 3.0 ± 0.5 6.5 ± 1.0 10.5 ± 1.0Construct the response matrix R using as basis functions 1, x,x 2 . Calculate thematrix R T N −1 R <strong>and</strong> show that its inverse, the covariance matrix V, hasthe<strong>for</strong>m⎛⎞V = 1 12 592 −9708 1580⎝−9708 8413 −1461⎠ .9184 1580 −1461 269Use this matrix to find the best values, <strong>and</strong> their uncertainties, <strong>for</strong> the coefficientsof the quadratic <strong>for</strong>m <strong>for</strong> y(x).31.17 The following are the values <strong>and</strong> st<strong>and</strong>ard errors of a physical quantity f(θ)measured at various values of θ (in which there is negligible error):θ 0 π/6 π/4 π/3f(θ) 3.72 ± 0.2 1.98 ± 0.1 −0.06 ± 0.1 −2.05 ± 0.1θ π/2 2π/3 3π/4 πf(θ) −2.83 ± 0.2 1.15 ± 0.1 3.99 ± 0.2 9.71 ± 0.4Theory suggests that f should be of the <strong>for</strong>m a 1 + a 2 cos θ + a 3 cos 2θ. Show thatthe normal equations <strong>for</strong> the coefficients a i are481.3a 1 + 158.4a 2 − 43.8a 3 = 284.7,158.4a 1 + 218.8a 2 +62.1a 3 = −31.1,−43.8a 1 +62.1a 2 + 131.3a 3 = 368.4.(a) If you have matrix inversion routines available on a computer, determine thebest values <strong>and</strong> variances <strong>for</strong> the coefficients a i <strong>and</strong> the correlation betweenthe coefficients a 1 <strong>and</strong> a 2 .(b) If you have only a calculator available, solve <strong>for</strong> the values using a Gauss–Seidel iteration <strong>and</strong> start from the approximate solution a 1 =2,a 2 = −2,a 3 =4.31.18 Prove that the expression given <strong>for</strong> the Student’s t-distribution in equation (31.118)is correctly normalised.31.19 Verify that the F-distribution P (F) given explicitly in equation (31.126) is symmetricbetween the two data samples, i.e. that it retains the same <strong>for</strong>m but with N 1<strong>and</strong> N 2 interchanged, if F is replaced by F ′ = F −1 . Symbolically, if P ′ (F ′ )isthedistribution of F ′ <strong>and</strong> P (F) =η(F,N 1 ,N 2 ), then P ′ (F ′ )=η(F ′ ,N 2 ,N 1 ).31.20 It is claimed that the two following sets of values were obtained (a) by r<strong>and</strong>omlydrawing from a normal distribution that is N(0, 1) <strong>and</strong> then (b) r<strong>and</strong>omlyassigning each reading to one of two sets A <strong>and</strong> B:Set A: −0.314 0.603 −0.551 −0.537 −0.160 −1.635 0.7190.610 0.482 −1.757 0.058Set B: −0.691 1.515 −1.642 −1.736 1.224 1.423 1.165Make tests, including t- <strong>and</strong>F-tests, to establish whether there is any evidencethat either claims is, or both claims are, false.1302


31.9 HINTS AND ANSWERS31.9 Hints <strong>and</strong> answers31.1 Note that the reading of 9.28 m s −2 is clearly in error, <strong>and</strong> should not be used inthe calculation; 9.80 ± 0.02 m s −2 .31.3 (a) 55.1. (b) Note that two thirds of the readings lie within ±2 of the mean <strong>and</strong>that 14 readings are being used. This gives a st<strong>and</strong>ard error in the mean ≈ 0.6.(c) Student’s t has a value of about 2.5 <strong>for</strong> 13 d.o.f. (degrees of freedom), <strong>and</strong>there<strong>for</strong>e it is likely at the 3% significance level that the data are in conflict withthe accepted value.31.5 Plot or calculate a least-squares fit of either x 2 versus x/y or xy versus y/xto obtain a ≈ 1.19 <strong>and</strong> b ≈ 3.4. (a) 0.16; (b) −0.27. Estimate (b) is the moreaccurate because, using the fact that y(−x) =−y(x), it is effectively obtained byinterpolation rather than extrapolation.31.7 Recall that, because of the equal proportions of each type, the expected numbersof each type in the first scheme is n. Show that the variance of the estimator <strong>for</strong>the second scheme is σ 2 /(kn). When calculating that <strong>for</strong> the first scheme, recallthat x 2 i= µ 2 i + σ 2 <strong>and</strong> note that µ 2 i can be written as (µ i − µ + µ) 2 .31.9 The log-likelihood function is()N∑N∑N∑ln L = ln n C xi + x i ln p + Nn − x i ln(1 − p);∂( n C x )∂ni=1≈ ln( nn − x)−i=1x2n(n − x) .Ignore the second term on the RHS of the above to obtainN∑( ) nln + N ln(1 − p) =0.n − x ii=131.11 ¯X =18.0 ± 2.2, Ȳ =15.0 ± 1.1. ˆσ =4.92 giving t =1.21 <strong>for</strong> 14 d.o.f., <strong>and</strong>is significant only at the 75% level. Thus there is no significant disagreementbetween the data <strong>and</strong> the theory. For the second theory, only the mean valuescan be tested as Y 2 will not be Gaussian distributed. The difference in the meansis Ȳ 2 − π 2 ¯X =47± 36 <strong>and</strong> is only significantly different from zero at the 82%level. Again the data is consistent with the proposed theory.31.13 Consider how many entries may be chosen freely in the table if all row <strong>and</strong>column totals are to match the observed values. It should be clear that <strong>for</strong> anm × n table the number of degrees of freedom is (m − 1)(n − 1).(a) In order to make the fractions expressing each preference or lack of preferencecorrect, the expected distribution, if there were no correlation, mustbeClassical None PopMathematics 17.5 10 22.5None 24.5 14 31.5English 28 16 36This gives a χ 2 of 12.3 <strong>for</strong> four d.o.f., making it less than 2% likely, that nocorrelation exists.(b) The expected distribution, if there were no correlation, isMusic preference No music preferenceAcademic preference 104 26No academic preference 56 141303i=1


STATISTICSThis gives a χ 2 of 1.2 <strong>for</strong> one d.o.f <strong>and</strong> no evidence <strong>for</strong> the claim.31.15 As the distribution at each value of s is Poisson, the best estimate of themeasurement error is the square root of the number of counts, i.e. √ n(s). Linearregression gives a =4.3 ± 2.1 <strong>and</strong>b =10.06 ± 0.94.(a) The cosmic ray background must be present, since n(0) ≠ 0 but its value ofabout 4 is uncertain to within a factor 2.(b) The correlation coefficient between a <strong>and</strong> b is −0.63. Yes; if a were reducedtowards zero then b would have to be increased to compensate.(c) Yes, χ 2 =4.9 <strong>for</strong> five d.o.f., which is almost exactly the ‘expected’ value,neither too good nor too bad.31.17 a 1 =2.02 ± 0.06, a 2 = −2.99 ± 0.09, a 3 =4.90 ± 0.10; r 12 = −0.60.31.19 Note that |dF| = |dF ′ /F ′2 | <strong>and</strong> write1+ N [ ][1 − 1N1 − 1as1+ (N ]2 − 1)F ′.(N 2 − 1)F ′ (N 2 − 1)F ′ N 1 − 11304


IndexWhere the discussion of a topic runs over two consecutive pages, reference is madeonly to the first of these. For discussions spread over three or more pages the first <strong>and</strong>last page numbers are given; these references are usually to the major treatment of thecorresponding topic. Isolated refences to a topic, including those appearing on consecutivepages, are listed individually. Some long topics are split, e.g. ‘Fourier trans<strong>for</strong>ms’ <strong>and</strong>‘Fourier trans<strong>for</strong>ms, examples’. The letter ‘n’ after a page number indicates that the topicis discussed in a footnote on the relevant page.A, B, one-dimensional irreps, 1090, 1102, 1109Abelian groups, 1044absolute convergence of series, 124, 831absolute derivative, 975–977acceleration vector, 335Adams method, 1024Adams–Moulton–Bash<strong>for</strong>th, predictor-correctorscheme, 1035addition rule <strong>for</strong> probabilities, 1125, 1130addition theorem <strong>for</strong> spherical harmonicsYl m (θ, φ), 594adjoint, see Hermitian conjugateadjoint operators, 559–564adjustment of parameters, 795Airy integrals, 890–894Ai(z), 889algebra ofcomplex numbers, 85functions in a vector space, 556matrices, 251power series, 134series, 131tensors, 938–941vectors, 213in a vector space, 242in component <strong>for</strong>m, 218algebraic equations, numerical methods <strong>for</strong>, seenumerical methods <strong>for</strong> algebraic equationsalternating group, 1116alternating series test, 130ammonia molecule, symmetries of, 1042Ampère’s rule (law), 381, 408amplitude modulation of radio waves, 444amplitude-phase diagram, 914analytic (regular) functions, 826angle between two vectors, 221angular frequency, 693nin Fourier series, 419angular momentum, 933, 949<strong>and</strong> irreps, 1093of particle system, 950–952of particles, 338of solid body, 396, 951vector representation, 238angular momentum operatorcomponent, 658total, 659angular velocity, vector representation, 223, 238,353annihilation <strong>and</strong> creation operators, 667anti-Hermitian matrices, 271eigenvalues, 276–278imaginary nature, 277eigenvectors, 276–278orthogonality, 277anti-Stokes line, 905anticommutativity of vector or cross product,222antisymmetric functions, 416<strong>and</strong> Fourier series, 419<strong>and</strong> Fourier trans<strong>for</strong>ms, 445antisymmetric matrices, 2701305


INDEXgeneral properties, see anti-Hermitianmatricesantisymmetric tensors, 938, 941antithetic variates, in Monte Carlo methods,1014aperture function, 437approximately equal ≈, definition, 132arbitrary parameters <strong>for</strong> ODE, 469arc length ofplane curves, 73space curves, 341arccosech, arccosh, arccoth, arcsech, arcsinh,arctanh, see hyperbolic functions, inversesArchimedean upthrust, 396, 410area element inCartesian coordinates, 188plane polars, 202area ofcircle, 71ellipse, 71, 207parallelogram, 223region, using multiple integrals, 191–193surfaces, 346as vector, 393–395, 408area, maximal enclosure, 779arg, argument of a complex number, 87Arg<strong>and</strong> diagram, 84, 825argument, principle of the, 880arithmetic series, 117arithmetico-geometric series, 118arrays, see matricesassociated Laguerre equation, 535, 621–624as example of Sturm–Liouville equation, 566,622natural interval, 567, 622associated Laguerre polynomials L m n (x), 621as special case of confluent hypergeometricfunction, 634generating function, 623orthogonality, 622recurrence relations, 624Rodrigues’ <strong>for</strong>mula, 622associated Legendre equation, 535, 587–593, 733,768general solution, 588as example of Sturm–Liouville equation, 566,590, 591general solution, 588natural interval, 567, 590, 591associated Legendre functions, 587–593of first kind Pl m (x), 588, 733, 768generating function, 592normalisation, 590orthogonality, 590, 591recurrence relations, 592Rodrigues’ <strong>for</strong>mula, 588of second kind Q m l (x), 588associative law <strong>for</strong>additionin a vector space of finite dimensionality,242in a vector space of infinite dimensionality,556of complex numbers, 86of matrices, 251of vectors, 213convolution, 447, 458group operations, 1043linear operators, 249multiplicationof a matrix by a scalar, 251ofavectorbyascalar,214of complex numbers, 88of matrices, 253multiplication by a scalarin a vector space of finite dimensionality,242in a vector space of infinite dimensionality,556atomic orbitals, 1115d-states, 1106, 1108, 1114p-states, 1106s-states, 1144auto-correlation functions, 450automorphism, 1061auxiliary equation, 493repeated roots, 493average value, see mean valueaxial vectors, 949backward differences, 1019basis functions<strong>for</strong> linear least squares estimation, 1273in a vector space of infinite dimensionality,556of a representation, 1078change in, 1084, 1087, 1092basis vectors, 217, 243, 929, 1078derivatives, 965–968Christoffel symbol Γ k ij , 965<strong>for</strong> particular irrep, 1106–1108, 1116linear dependence <strong>and</strong> independence, 217non-orthogonal, 245orthonormal, 244required properties, 217Bayes’ theorem, 1132Bernoulli equation, 477Bessel correction to variance estimate, 1248Bessel equation, 535, 602–607, 614, 615as example of Sturm–Liouville equation, 566natural interval, 608Bessel functions J ν (x), 602–614, 729, 738as special case of confluent hypergeometricfunction, 634generating function, 613graph of, 606integral relationships, 610integral representation, 613–614orthogonality, 608–6111306


INDEXrecurrence relations, 611–612second kind Y ν (x), 607graph of, 607series, 604ν = 0, 606ν = ±1/2, 605spherical j l (x), 615, 741zeros of, 729, 739Bessel inequality, 246, 559best unbiased estimator, 1232beta function, 638bias of estimator, 1231bilinear trans<strong>for</strong>mation, general, 110binary chopping, 990binomial coefficient n C k , 27–30, 1135–1137elementary properties, 26identities, 27in Leibnitz’ theorem, 49negative n, 29non-integral n, 29binomial distribution Bin(n, p), 1168–1171<strong>and</strong> Gaussian distribution, 1185<strong>and</strong> Poisson distribution, 1174, 1177mean <strong>and</strong> variance, 1171MGF, 1170recurrence <strong>for</strong>mula, 1169binomial expansion, 25–30, 140binormal to space curves, 342birthdays, different, 1134bivariate distributions, 1196–1207conditional, 1198continuous, 1197correlation, 1200–1207<strong>and</strong> independence, 1200matrix, 1203–1207positive <strong>and</strong> negative, 1200uncorrelated, 1200covariance, 1200–1207matrix, 1203expectation (mean), 1199independent, 1197, 1200marginal, 1198variance, 1200Boltzmann distribution, 171bonding in molecules, 1103, 1105–1108Born approximation, 149, 575Bose–Einstein statistics, 1138boundary conditions<strong>and</strong> characteristics, 700<strong>and</strong> Laplace equation, 764, 766<strong>for</strong> Green’s functions, 512, 514–516inhomogeneous, 515<strong>for</strong> ODE, 468, 470, 501<strong>for</strong> PDE, 681, 685–687<strong>for</strong> Sturm–Liouville equations, 564homogeneous <strong>and</strong> inhomogeneous, 685, 723,752, 754superposition solutions, 718–724types, 702–705bra vector 〈ψ|, 649brachistochrone problem, 784Bragg <strong>for</strong>mula, 237branch cut, 835branch points, 835Bromwich integral, 884bulk modulus, 980calculus of residues, see zeros of a function of acomplex variable <strong>and</strong> contour integrationcalculus of variationsconstrained variation, 785–787estimation of ODE eigenvalues, 790Euler–Lagrange equation, 776Fermat’s principle, 787Hamilton’s principle, 788higher-order derivatives, 782several dependent variables, 782several independent variables, 782soap films, 780variable end-points, 782–785calculus, elementary, 41–76cancellation law in a group, 1046canonical <strong>for</strong>m, <strong>for</strong> second-order ODE, 516card drawing, see probabilitycarrier frequency of radio waves, 445Cartesian coordinates, 217Cartesian tensors, 930–955algebra, 938–941contraction, 939definition, 935first-order, 932–935from scalar, 934general order, 935–954integral theorems, 954isotropic, 944–946physical applications, 934, 939–941, 950–954second-order, 935–954, 968symmetry <strong>and</strong> antisymmetry, 938tensor fields, 954zero-order, 932–935from vector, 935Cartesian tensors, particularconductivity, 952inertia, 951strain, 953stress, 953susceptibility, 952catenary, 781, 787Cauchyboundary conditions, 702distribution, 1152inequality, 853integrals, 851–853product, 131root test, 129, 831theorem, 849Cauchy–Riemann relations, 827–830, 849, 873,875in terms of z <strong>and</strong> z ∗ , 829central differences, 10191307


INDEXcentral limit theorem, 1194–1196central moments, see moments, centralcentre of a group, 1069centre of mass, 195of hemisphere, 195of semicircular lamina, 197centroid, 195of plane area, 195of plane curve, 197of triangle, 216CF, see complementary functionchain rule <strong>for</strong> functions ofone real variable, 46several real variables, 157change of basis, see similarity trans<strong>for</strong>mationschange of variables<strong>and</strong> coordinate systems, 158–160in multiple integrals, 199–207evaluation of Gaussian integral, 202–204general properties, 206in RVD, 1150–1157character tables, 10933m or 32 or C 3v or S 3 , 1093, 1097, 1108, 1110,11174mm or C 4v or D 4 , 1102, 1106, 1108, 1113A 4 , 1116D 5 , 1116S 4 or 432 or O, 1114, 1115¯43m or T d , 1115construction of, 1100–1102quaternion, 1113characteristic equation, 280normal mode <strong>for</strong>m, 319of recurrence relation, 499characteristic functions, see moment generatingfunctions (MGFs)characteristics<strong>and</strong> boundary curves, 700multiple intersections, 700, 705<strong>and</strong> the existence of solutions, 699–705first-order equations, 699second-order equations, 703<strong>and</strong> equation type, 703characters, 1092–1096, 1100–1102<strong>and</strong> conjugacy classes, 1092, 1095counting irreps, 1095definition, 1092of product representation, 1104orthogonality properties, 1094, 1102summation rules, 1097charge (point), Dirac δ-function respresentation,441charged particle in electromagnetic fields, 370Chebyshev equation, 535, 595–602as example of Sturm–Liouville equation, 566,599general solution, 597natural interval, 567, 599polynomial solutions, 552Chebyshev functions, 595–602Chebyshev polynomialsof first kind T n (x), 596as special case of hypergeometric function,631generating function, 601graph of, 597normalisation, 600orthogonality, 599recurrence relations, 601Rodrigues’ <strong>for</strong>mula, 599of second kind U n (x), 597generating function, 601graph of, 598normalisation, 600orthogonality, 599Rodrigues’ <strong>for</strong>mula, 599chi-squared (χ 2 ) distribution, 1192<strong>and</strong> goodness of fit, 1297<strong>and</strong> likelihood-ratio test, 1283, 1292<strong>and</strong> multiple estimators, 1243percentage points tabulation, 1244test <strong>for</strong> correlation, 1301Cholesky separation, 313Christoffel symbol Γ k ij , 965–968from metric tensor, 966, 973circlearea of, 71equation <strong>for</strong>, 16circle of convergence, 831Clairaut equation, 483classes <strong>and</strong> equivalence relations, 1064closure of a group, 1043closure property of eigenfunctions of anHermitian operator, 563cofactor of a matrix element, 259column matrix, 250column vector, 250combinations (probability), 1133–1139common ratio in geometric series, 117commutation law <strong>for</strong> group elements, 1044commutative law <strong>for</strong>additionin a vector space of finite dimensionality,242in a vector space of infinite dimensionality,556of complex numbers, 86of matrices, 251of vectors, 213complex scalar or dot product, 222convolution, 447, 458inner product, 244multiplicationofavectorbyascalar,214of complex numbers, 88scalar or dot product, 220commutatorof two matrices, 309of two operators, 653, 656comparison test, 1251308


INDEXcomplement, 1121probability <strong>for</strong>, 1125complementary equation, 490complementary error function, 640complementary function (CF), 491<strong>for</strong> ODE, 492partially known, 506repeated roots of auxiliary equation, 493completeness ofbasis vectors, 243eigenfunctions of an Hermitian operator, 560,563eigenvectors of a normal matrix, 275spherical harmonics Yl m (θ, φ), 594completing the squareas a means of integration, 66<strong>for</strong> quadratic equations, 35<strong>for</strong> quadratic <strong>for</strong>ms, 1206to evaluate Gaussian integral, 436, 749complex conjugatez ∗ , of complex number, 89–91, 829of a matrix, 256–258of scalar or dot product, 222properties of, 90complex exponential function, 92, 833complex Fourier series, 424complex integrals, 845–849, see also zeros of afunction of a complex variable <strong>and</strong> contourintegrationAiry integrals, 890–894Cauchy integrals, 851–853Cauchy’s theorem, 849definition, 845Jordan’s lemma, 864Morera’s theorem, 851of z −1 , 846principal value, 864residue theorem, 858–860WKB methods, 895–905complex logarithms, 99, 834principal value of, 100, 834complex numbers, 83–114addition <strong>and</strong> subtraction of, 85applications to differentiation <strong>and</strong> integration,101argument of, 87associativity ofaddition, 86multiplication, 88commutativity ofaddition, 86multiplication, 88complex conjugate of, see complex conjugatecomponents of, 84de Moivre’s theorem, see de Moivre’s theoremdivision of, 91, 94from roots of polynomial equations, 83imaginary part of, 83modulus of, 87multiplication of, 88, 94as rotation in the Arg<strong>and</strong> diagram, 88notation, 84polar representation of, 92–95real part of, 83trigonometric representation of, 93complex potentials, 871–876<strong>and</strong> fluid flow, 873equipotentials <strong>and</strong> field lines, 872<strong>for</strong> circular <strong>and</strong> elliptic cylinders, 876<strong>for</strong> parallel cylinders, 921<strong>for</strong> plates, 877–879, 921<strong>for</strong> strip, 921<strong>for</strong> wedges, 878under con<strong>for</strong>mal trans<strong>for</strong>mations, 876–879complex power series, 133complex powers, 99complex variables, see functions of a complexvariable <strong>and</strong> power series in a complexvariable <strong>and</strong> complex integralscomponentsof a complex number, 84of a vector, 217in a non-orthogonal basis, 234uniqueness, 243conditional (constrained) variation, 785–787conditional convergence, 124conditional distributions, 1198conditional probability, see probability,conditionalconesurface area of, 74volume of, 75confidence interval, 1236confidence region, 1241confluence process, 634confluent hypergeometric equation, 535, 633as example of Sturm–Liouville equation, 566general solution, 633confluent hypergeometric functions, 633contiguous relations, 635integral representation, 634recurrence relations, 635special cases, 634con<strong>for</strong>mal trans<strong>for</strong>mations (mappings), 839–879applications, 876–879examples, 842–844properties, 839–842Schwarz–Christoffel trans<strong>for</strong>mation, 843congruence, 1065conic sections, 15eccentricity, 17parametric <strong>for</strong>ms, 17st<strong>and</strong>ard <strong>for</strong>ms, 16conjugacy classes, 1068–1070element in a class by itself, 1068conjugate roots of polynomial equations, 99connectivity of regions, 383conservative fields, 387–389necessary <strong>and</strong> sufficient conditions, 387–389potential (function), 3891309


INDEXconsistency, of estimator, 1230constant coefficients in ODE, 492–503auxiliary equation, 493constants of integration, 62, 468constrained variation, 785–787constraints, stationary values under, seeLagrange undetermined multiplerscontinuity correction <strong>for</strong> discrete RV, 1186continuity equation, 404contour integration, 861–867, 887infinite integrals, 862–867inverse Laplace trans<strong>for</strong>ms, 884–887residue theorem, 858–867sinusoidal functions, 861summing series, 882contraction of tensors, 939contradiction, proof by, 32–34contravariantbasis vectors, 961derivative, 965components of tensor, 956definition, 961control variates, in Monte Carlo methods, 1013convergence of infinite series, 831absolute, 124, 831complex power series, 133conditional, 124necessary condition, 125power series, 132under various manipulations, see powerseries, manipulationratio test, 832rearrangement of terms, 124tests <strong>for</strong> convergence, 125–131alternating series test, 130comparison test, 125grouping terms, 129integral test, 128quotient test, 127ratio comparison test, 127ratio test (D’Alembert), 126, 132root test (Cauchy), 129, 831convergence of numerical iteration schemes,992–994convolutionFourier tran<strong>for</strong>ms, see Fourier trans<strong>for</strong>ms,convolutionLaplace tran<strong>for</strong>ms, see Laplace trans<strong>for</strong>ms,convolutionconvolution theoremFourier trans<strong>for</strong>ms, 448Laplace trans<strong>for</strong>ms, 457coordinate geometry, 15–18conic sections, 15straight line, 15coordinate systems, see Cartesian, curvilinear,cylindrical polar, plane polar <strong>and</strong> sphericalpolar coordinatescoordinate trans<strong>for</strong>mations<strong>and</strong> integrals, see change of variables<strong>and</strong> matrices, see similarity trans<strong>for</strong>mationsgeneral, 960–965relative tensors, 963tensor trans<strong>for</strong>mations, 962weight, 964orthogonal, 932coplanar vectors, 225Cornu spiral, 914correlation functions, 449–451auto-correlation, 450cross-correlation, 449energy spectrum, 450Parseval’s theorem, 451Wiener–Kinchin theorem, 450correlation matrix, of sample, 1229correlation of bivariate distributions, 1200–1207correlation of sample data, 1229correlation, chi-squared test, 1301correspondence principle in quantum mechanics,1215cosets <strong>and</strong> congruence, 1065cosh, hyperbolic cosine, 102, 833, see alsohyperbolic functionscosine, cos(x)in terms of exponential functions, 102Maclaurin series <strong>for</strong>, 140orthogonality relations, 417counting irreps, see characters, counting irrepscoupled pendulums, 329, 331covariance matrixof linear least squares estimators, 1274of sample, 1229covariance of bivariate distributions, 1200–1207covariance of sample data, 1229covariantbasis vector, 961derivative, 965components of tensor, 956definition, 961derivative, 968of scalar, 971semi-colon notation, 969differentiation, 968–971CPF, see probability functions, cumulativeCramér–Rao (Fisher’s) inequality, 1232, 1233Cramer determinant, 299Cramer’s rule, 299cross product, see vector productcross-correlation functions, 449crystal lattice, 148crystal point groups, 1082cube roots of unity, 98cube, rotational symmetries of, 1114cumulants, 1166curl of a vector field, 353as a determinant, 353as integral, 398, 400curl curl, 356in curvilinear coordinates, 368in cylindrical polars, 3601310


INDEXin spherical polars, 362Stoke’s theorem, 406–409tensor <strong>for</strong>m, 974current-carrying wire, magnetic potential, 729curvature, 52–55circle of, 53of a function, 52of space curves, 342radius of, 53curves, see plane curves <strong>and</strong> space curvescurvilinear coordinates, 364–369basis vectors, 364length <strong>and</strong> volume elements, 365scale factors, 364surfaces <strong>and</strong> curves, 364tensors, 955–977vector operators, 367–369cut plane, 865cycle notation <strong>for</strong> permutations, 1057cyclic groups, 1061, 1098cyclic relation <strong>for</strong> partial derivatives, 157cycloid, 370, 785cylinders, conducting, 874, 876cylindrical polar coordinates, 357–361area element, 360basis vectors, 358Laplace equation, 728–731length element, 360vector operators, 357–361volume element, 360δ-function (Dirac), see Dirac δ-functionδ ij , δ j i , Kronecker delta, tensor, see Kroneckerdelta, δ ij , δ j i ,tensorD’Alembert’s ratio test, 126, 832in convergence of power series, 132D’Alembert’s solution to wave equation, 694damped harmonic oscillators, 239<strong>and</strong> Parseval’s theorem, 451data modelling, maximum-likelihood, 1255de Broglie relation, 436, 709, 768de Moivre’s theorem, 95, 861applications, 95–99finding the nth roots of unity, 97solving polynomial equations, 98trigonometric identities, 95–97deconvolution, 449defective matrices, 278, 311degeneracybreaking of, 1111–1113of normal modes, 1110degenerate (separable) kernel, 807degenerate eigenvalues, 275, 282degreeof ODE, 468of polynomial equation, 2del ∇, see gradient operator (grad)del squared ∇ 2 (Laplacian), 352, 676as integral, 400in curvilinear coordinates, 368in cylindrical polar coordinates, 360in polar coordinates, 725in spherical polar coordinates, 362, 741tensor <strong>for</strong>m, 973delta function (Dirac), see Dirac δ-functiondependent r<strong>and</strong>om variables, 1196–1205derivative, see also differentiationabsolute, 975–977covariant, 968Fourier trans<strong>for</strong>m of, 444Laplace trans<strong>for</strong>m of, 455normal, 350of basis vectors, 336of composite vector expressions, 337of function of a complex variable, 825of function of a function, 46of hyperbolic functions, 106–109of products, 44–46, 48–50of quotients, 47of simple functions, 44of vectors, 334ordinary, first, second <strong>and</strong> nth, 42partial, see partial differentiationtotal, 154derivative method <strong>for</strong> second series solution ofODE, 545–548determinant <strong>for</strong>m<strong>and</strong> ɛ ijk , 942<strong>for</strong> curl, 353determinants, 259–263adding rows or columns, 262<strong>and</strong> singular matrices, 263as product of eigenvalues, 287evaluationusing ɛ ijk , 942using Laplace expansion, 259identical rows or columns, 262in terms of cofactors, 259interchanging two rows or two columns, 262Jacobian representation, 201, 205, 207notation, 259of Hermitian conjugate matrices, 262of order three, in components, 260of transpose matrices, 261product rule, 262properties, 261–263, 978relationship with rank, 267removing factors, 262secular, 280diagonal matrices, 268diagonalisation of matrices, 285–288normal matrices, 286properties of eigenvalues, 287simultaneous, 331diamond, unit cell, 234die throwing, see probabilitydifference method <strong>for</strong> summation of series, 119difference schemes <strong>for</strong> differential equations,1020–1023, 1030–10321311


INDEXdifference, finite, see finite differencesdifferentiablefunction of a complex variable, 825–827function of a real variable, 42differentialdefinition, 43exact <strong>and</strong> inexact, 155of vector, 338, 344total, 154differential equations, see ordinary differentialequations <strong>and</strong> partial differential equationsdifferential equations, particularassociated Laguerre, 535, 566, 621–624associated Legendre, 535, 566, 587–593Bernoulli, 477Bessel, 535, 566, 602–607, 614Chebyshev, 535, 566, 595–602Clairaut, 483confluent hypergeometric, 535, 566, 633diffusion, 678, 695–698, 716, 723, 1032Euler, 504Euler–Lagrange, 776Helmholtz, 737–741Hermite, 535, 566, 624–628hypergeometric, 535, 566, 628–632Lagrange, 789Laguerre, 535, 566, 616–621Laplace, 679, 690, 717, 718, 1031Legendre, 534, 535, 566, 577–586Legendre linear, 503–505Poisson, 679, 744–746Schrödinger, 679, 741, 768, 795simple harmonic oscillator, 535, 566Sturm–Liouville, 790wave, 676, 689, 693–695, 714, 737, 790differential operators, see linear differentialoperatordifferentiation, see also derivativeas gradient, 42as rate of change, 41chain rule, 46covariant, 968–971from first principles, 41–44implicit, 47logarithmic, 48notation, 43of Fourier series, 424of integrals, 178of power series, 135partial, see partial differentiationproduct rule, 44–46, 48–50quotient rule, 47theorems, 55–57using complex numbers, 101diffraction, see Fraunhofer diffractiondiffusion equation, 678, 688, 695–698combination of variables, 696–698integral trans<strong>for</strong>ms, 747numerical methods, 1032separation of variables, 716simple solution, 696superposition, 723diffusion of solute, 678, 696, 747dihedral group, 1113, 1116dimension of irrep, 1088dimensionality of vector space, 243dipole matrix elements, 208, 1108, 1115dipole moments of molecules, 1077Dirac δ-function, 355, 405, 439–443<strong>and</strong> convolution, 447<strong>and</strong> Green’s functions, 511, 512as limit of various distributions, 443as sum of harmonic waves, 442definition, 439Fourier trans<strong>for</strong>m of, 443impulses, 441point charges, 441properties, 439reality of, 443relation to Fourier trans<strong>for</strong>ms, 442relation to Heaviside (unit step) function, 441three-dimensional, 441, 452Dirac notation, 648direct product, of groups, 1072direct sum ⊕, 1086direction cosines, 221Dirichlet boundary conditions, 702, 852nGreen’s functions, 754, 756–765method of images, 758–765Dirichlet conditions, <strong>for</strong> Fourier series, 415disc, moment of inertia, 208discontinuous functions <strong>and</strong> Fourier series,420–422discrete Fourier trans<strong>for</strong>ms, 462disjoint events, see mutually exclusive eventsdisplacement kernel, 809distance from aline to a line, 231line to a plane, 232point to a line, 229point to a plane, 230distributive law <strong>for</strong>addition of matrix products, 254convolution, 447, 458inner product, 244linear operators, 249multiplicationof a matrix by a scalar, 251of a vector by a complex scalar, 222ofavectorbyascalar,214multiplication by a scalarin a vector space of finite dimensionality,242in a vector space of infinite dimensionality,556scalar or dot product, 220vector or cross product, 222div, divergence of vector fields, 352as integral, 398in curvilinear coordinates, 3671312


INDEXin cylindrical polars, 360in spherical polars, 362tensor <strong>for</strong>m, 972divergence theorem<strong>for</strong> tensors, 954<strong>for</strong> vectors, 401in two dimensions, 384physical applications, 404related theorems, 403division axiom in a group, 1046division of complex numbers, 91dominant term, in Stokes phenomenon, 904dot product, see scalar productdouble integrals, see multiple integralsdrumskin, see membranedual tensors, 949dummy variable, 61ɛ ijk , Levi-Civita symbol, tensor, 941–946<strong>and</strong> determinant, 942identities, 943isotropic, 945vector products, 942weight, 964e x , see exponential functionE, two-dimensional irrep, 1090, 1102, 1108eccentricity, of conic sections, 17efficiency, of estimator, 1231eigenequation <strong>for</strong> differential operators, 554more general <strong>for</strong>m, 555, 571–573eigenfrequencies, 319estimation using Rayleigh–Ritz method,327–329eigenfunctionscompleteness <strong>for</strong> Hermitian operators, 560,563construction of a real set <strong>for</strong> an Hermitianoperator, 563definition, 555normalisation <strong>for</strong> Hermitian operators, 562of integral equations, 817of simple harmonic oscillators, 555orthogonality <strong>for</strong> Hermitian operators,561–563eigenvalues, 272–282, see Hermitian operatorscharacteristic equation, 280continuous <strong>and</strong> discrete, 650definition, 272degenerate, 282determination, 280–282estimation <strong>for</strong> ODE, 790estimation using Rayleigh–Ritz method,327–329notation, 273of anti-Hermitian matrices, seeanti-Hermitian matricesof Fredholm equations, 808of general square matrices, 278of Hermitian matrices, see Hermitian matricesof integral equations, 808, 816of linear differential operatorsadjustment of parameters, 795definition, 555error in estimate of, 793estimation, 790–796higher eigenvalues, 793, 800simple harmonic oscillator, 555of linear operators, 272of normal matrices, 273–276of representative matrices, 1100of unitary matrices, 278under similarity trans<strong>for</strong>mation, 287eigenvectors, 272–282characteristic equation, 280definition, 272determination, 280–282normalisation condition, 273notation, 273of anti-Hermitian matrices, seeanti-Hermitian matricesof commuting matrices, 278of general square matrices, 278of Hermitian matrices, see Hermitian matricesof linear operators, 272of normal matrices, 273–276of unitary matrices, 278stationary properties <strong>for</strong> quadratic <strong>and</strong>Hermitian <strong>for</strong>ms, 290Einstein relation, 436, 709, 768elastic de<strong>for</strong>mations, 953electromagnetic fieldsflux, 395Maxwell’s equations, 373, 408, 979electrostatic fields <strong>and</strong> potentialscharged split sphere, 735conducting cylinder in uni<strong>for</strong>m field, 876conducting sphere in uni<strong>for</strong>m field, 734from charge density, 745, 758from complex potential, 873infinite charged plate, 759, 877infinite wedge with line charge, 878ininite charged wedge, 877of line charges, 761, 872semi-infinite charged plate, 877sphere with point charge, 764ellipsearea of, 71, 207, 385as section of quadratic surface, 292equation <strong>for</strong>, 16ellipsoid, volume of, 207elliptic PDE, 687, 690empty event ∅, 1121end-points <strong>for</strong> variationscontributions from, 782fixed, 777variable, 782–785energy levels ofparticle in a box, 768simple harmonic oscillator, 6421313


INDEXenergy spectrum <strong>and</strong> Fourier trans<strong>for</strong>ms, 450,451entire functions, 832nenvelopes, 173–175equations of, 174to a family of curves, 173epimorphism, 1061equilateral triangle, symmetries of, 1047, 1052,1081, 1110equivalence relations, 1064–1066, 1068<strong>and</strong> classes, 1064congruence, 1065–1067examples, 1070equivalence trans<strong>for</strong>mations, see similaritytrans<strong>for</strong>mationsequivalent representations, 1084–1086, 1099error function, erf(x), 640, 697, 748as special case of confluent hypergeometricfunction, 634error termsin Fourier series, 430in Taylor series, 139errors, first <strong>and</strong> second kind, 1280essential singularity, 838, 856estimation of eigenvalueslinear differential operator, 792–795Rayleigh–Ritz method, 327–329estimators (statistics), 1229best unbiased, 1232bias, 1231central confidence interval, 1237confidence interval, 1236confidence limits, 1236confidence region, 1241consistency, 1230efficiency, 1231maximum-likelihood, 1256minimum-variance, 1232st<strong>and</strong>ard error, 1234Euler equationdifferential, 504, 522trigonometric, 93Euler method, numerical, 1021Euler–Lagrange equation, 776special cases, 777–781even functions, see symmetric functionsevents, 1120complement of, 1121empty ∅, 1121intersection of ∩, 1120mutually exclusive, 1129statistically independent, 1129union of ∪, 1121exact differentials, 155exact equations, 472, 505condition <strong>for</strong>, 472non-linear, 519expectation values, see probability distributions,meanexponential distribution, 1190from Poisson, 1190MGF, 1191exponential functionMaclaurin series <strong>for</strong>, 140of a complex variable, 92, 833relation with hyperbolic functions, 102F-distribution (Fisher), 1290–1296critical points table, 1295logarithmic <strong>for</strong>m, 1296Fabry–Pérot interferometer, 146factorial function, general, 636factorisation, of a polynomial equation, 7faithful representation, 1083, 1098Fermat’s principle, 787, 798Fermi–Dirac statistics, 1138Fibonacci series, 525field lines <strong>and</strong> complex potentials, 872fieldsconservative, 387–389scalar, 347tensor, 954vector, 347fields, electrostatic, see electrostatic fields <strong>and</strong>potentialsfields, gravitational, see gravitational fields <strong>and</strong>potentialsfinite differences, 1019central, 1019<strong>for</strong> differential equations, 1020–1023<strong>for</strong>ward <strong>and</strong> backward, 1019from Taylor series, 1019, 1026schemes <strong>for</strong> differential equations, 1030–1032finite groups, 1043first law of thermodynamics, 176first-order differential equations, see ordinarydifferential equationsFisher distribution, see F-distribution (Fisher)Fisher matrix, 1241, 1268Fisher’s inequality, 1232, 1233fluidsArchimedean upthrust, 396, 410complex velocity potential, 873continuity equation, 404cylinder in uni<strong>for</strong>m flow, 874flow, 873flux, 395, 875irrotational flow, 353sources <strong>and</strong> sinks, 404, 873stagnation points, 873velocity potential, 409, 679vortex flow, 408, 874<strong>for</strong>ward differences, 1019Fourier cosine trans<strong>for</strong>ms, 446Fourier series, 415–432<strong>and</strong> separation of variables, 719–722, 724coefficients, 417–419, 425complex, 424differentiation, 424Dirichlet conditions, 4151314


INDEXdiscontinuous functions, 420–422error term, 430integration, 424non-periodic functions, 422–424orthogonality of terms, 417complex case, 425Parseval’s theorem, 426st<strong>and</strong>ard <strong>for</strong>m, 417summation of series, 427symmetry considerations, 419uses, 415Fourier series, examplessquare-wave, 418x, 424, 425x 2 , 422x 3 , 424Fourier sine trans<strong>for</strong>ms, 445Fourier trans<strong>for</strong>ms, 433–453as generalisation of Fourier series, 433–435convolution, 446–449<strong>and</strong> the Dirac δ-function, 447associativity, commutativity, distributivity,447definition, 447resolution function, 446convolution theorem, 448correlation functions, 449–451cosine trans<strong>for</strong>ms, 446deconvolution, 449definition, 435discrete, 462evaluation using convolution theorem, 448<strong>for</strong> integral equations, 809–812<strong>for</strong> PDE, 749–751Fourier-related (conjugate) variables, 436in higher dimensions, 451–453inverse, definition, 435odd <strong>and</strong> even functions, 445Parseval’s theorem, 450properties: differentiation, exponentialmultiplication, integration, scaling,translation, 444relation to Dirac δ-function, 442sine trans<strong>for</strong>ms, 445Fourier trans<strong>for</strong>ms, examplesconvolution, 448damped harmonic oscillator, 451Dirac δ-function, 443exponential decay function, 435Gaussian (normal) distribution, 435rectangular distribution, 442spherically symmetric functions, 452two narrow slits, 448two wide slits, 438, 448Fourier’s inversion theorem, 435Fraunhofer diffraction, 437–439diffraction grating, 461two narrow slits, 448two wide slits, 438, 448Fredholm integral equations, 805eigenvalues, 808operator <strong>for</strong>m, 806with separable kernel, 807Fredholm theory, 815Frenet–Serret <strong>for</strong>mulae, 343Fresnel integrals, 913Frobenius series, 539Fuch’s theorem, 539function of a matrix, 255functional, 776functions of a complex variable, 825–839,853–858analyticity, 826behaviour at infinity, 839branch points, 835Cauchy integrals, 851–853Cauchy–Riemann relations, 827–830con<strong>for</strong>mal trans<strong>for</strong>mations, 839–844derivative, 825differentiation, 825–830identity theorem, 854Laplace equation, 829, 871Laurent expansion, 855–858multivalued <strong>and</strong> branch cuts, 835–837, 885particular functions, 832–835poles, 837power series, 830–832real <strong>and</strong> imaginary parts, 825, 830singularities, 826, 837–839Taylor expansion, 853–855zeros, 839, 879–882functions of one real variabledecomposition into even <strong>and</strong> odd functions,416differentiation of, 41–50Fourier series, see Fourier seriesintegration of, 59–72limits, see limitsmaxima <strong>and</strong> minima of, 50–52stationary values of, 50–52Taylor series, see Taylor seriesfunctions of several real variableschain rule, 157differentiation of, 151–179integration of, see multiple integrals,evaluationmaxima <strong>and</strong> minima, 162–167points of inflection, 162–167rates of change, 153–155saddle points, 162–167stationary values, 162–167Taylor series, 160–162fundamental solution, 757fundamental theorem ofalgebra, 83, 85, 868calculus, 61complex numbers, see de Moivre’s theoremgamma distribution, 1153, 1191gamma function1315


INDEXas general factorial function, 636definition <strong>and</strong> properties, 636graph of, 637Gauss’s theorem, 765Gauss–Seidel iteration, 996–998Gaussian (normal) distribution N(µ, σ 2 ),1179–1189<strong>and</strong> binomial distribution, 1185<strong>and</strong> central limit theorem, 1195<strong>and</strong> Poisson distribution, 1187continuity correction, 1186CPF, 1018, 1181tabulation, 1182Fourier trans<strong>for</strong>m, 435integration with infinite limits, 202–204mean <strong>and</strong> variance, 1180–1184MGF, 1185, 1188multiple, 1188multivariate, 1209r<strong>and</strong>om number generation, 1018sigma limits, 1183st<strong>and</strong>ard variable, 1180Gaussian elimination with interchange, 995Gaussian integration, 1005–1009points <strong>and</strong> weights, 1008, 1010general tensorsalgebra, 938–941contraction, 939contravariant, 961covariant, 961dual, 949metric, 957–960physical applications, 957–960, 976pseudotensors, 964tensor densities, 964generalised likelihood ratio, 1282generating functionsassociated Laguerre polynomials, 623associated Legendre polynomials, 592Bessel functions, 613Chebyshev polynomials, 601Hermite polynomials, 627Laguerre polynomials, 620Legendre polynomials, 584–586generating functions in probability, 1157–1167,see also moment generating functions <strong>and</strong>probability generating functionsgeodesics, 797, 976, 982geometric distribution, 1159, 1172geometric series, 117Gibbs’ free energy, 178Gibbs’ phenonmenon, 421gradient of a function ofone variable, 42several real variables, 153–155gradient of scalar, 348–352tensor <strong>for</strong>m, 972gradient of vector, 937, 969gradient operator (grad), 348as integral, 398in curvilinear coordinates, 367in cylindrical polars, 360in spherical polars, 362tensor <strong>for</strong>m, 972Gram–Schmidt orthogonalisation ofeigenfunctions of Hermitian operators, 562eigenvectors ofHermitian matrices, 277normal matrices, 275functions in a Hilbert space, 557gravitational fields <strong>and</strong> potentialsLaplace equation, 679Newton’s law, 339Poisson equation, 679, 744uni<strong>for</strong>m disc, 771uni<strong>for</strong>m ring, 742Green’s functions, 568–571, 751–767<strong>and</strong> boundary conditions, 512, 514<strong>and</strong> Dirac δ-function, 511<strong>and</strong> partial differential operators, 753<strong>and</strong> Wronskian, 527diffusion equation, 749Dirichlet problems, 756–765<strong>for</strong> ODE, 185, 511–516Neumann problems, 765–767particular integrals from, 514Poisson’s equation, 755Green’s theoremsapplications, 706, 754, 849in a plane, 384–387, 407in three dimensions, 402ground-state energyharmonic oscillator, 796hydrogen atom, 800group multiplication tables, 1050order three, 1062order four, 1050, 1052, 1061order five, 1062order six, 1055, 1061grouping terms as a test <strong>for</strong> convergence, 129groupsAbelian, 1044associative law, 1043cancellation law, 1046centre, 1069closure, 1043cyclic, 1061definition, 1043–1046direct product, 1072division axiom, 1046elements, 1043order, 1047finite, 1043identity element, 1043–1046inverse, 1043, 1046isomorphic, 1051mappings between, 1059–1061homomorphic, 1059–1061image, 1059isomorphic, 10591316


INDEXnomenclature, 1102non-Abelian, 1052–1056order, 1043, 1081, 1082, 1094, 1097, 1100permutation law, 1047subgroups, see subgroupsgroups, examples1<strong>and</strong>−1 under multiplication, 1043alternating, 1116complex numbers e iθ , 1048functions, 1055general linear, 1073integers under addition, 1043integers under multiplication (mod N),1049–1051matrices, 1054permutations, 1056–1058quaternion, 1073rotation matrices, 1048symmetries of a square, 1100symmetries of an equilateral triangle, 1047H n (x), see Hermite polynomialsHamilton’s principle, 788Hamiltonian, 796Hankel functions H ν (1) (x), H ν (2) (x), 607Hankel trans<strong>for</strong>ms, 459harmonic oscillatorsdamped, 239, 451ground-state energy, 796Schrödinger equation, 796simple, see simple harmonic oscillatorheat flowdiffusion equation, 678, 696, 723in bar, 723, 749, 770in thin sheet, 698Heaviside function, 441relation to Dirac δ-function, 441Heisenberg’s uncertainty principle, 435–437Helmholtz equation, 737–741cylindrical polars, 740plane polars, 738spherical polars, 740–741Helmholtz potential, 177hemisphere, centre of mass <strong>and</strong> centroid, 195Hermite equation, 535, 624–628as example of Sturm–Liouville equation, 566natural interval, 567Hermite polynomials H n (x), 625as special case of confluent hypergeometricfunction, 634generating function, 627graph of, 625normalisation, 626orthogonality, 626recurrence relations, 628Rodrigues’ <strong>for</strong>mula, 626Hermitian conjugate, 256–258<strong>and</strong> inner product, 258product rule, 257Hermitian <strong>for</strong>ms, 288–292positive definite <strong>and</strong> semi-definite, 290stationary properties of eigenvectors, 290Hermitian kernel, 816Hermitian matrices, 271eigenvalues, 276–278reality, 276eigenvectors, 276–278orthogonality, 277Hermitian operators, 559–564<strong>and</strong> physical variables, 650boundary condition <strong>for</strong> simple harmonicoscillators, 560eigenfunctionscompleteness, 560, 563orthogonality, 561–563eigenvaluesreality, 561Green’s functions, 568–571importance of, 555, 560in Sturm–Liouville equations, 564properties, 561–564superposition methods, 568–571higher-order differential equations, see ordinarydifferential equationsHilbert spaces, 557–559hit or miss, in Monte Carlo methods, 1014homogeneousboundary conditions, see boundaryconditions, homogeneous <strong>and</strong>inhomogeneousdifferential equations, 490dimensionally consistent, 475, 521simultaneous linear equations, 293homomorphism, 1059–1061kernel of, 1060representation as, 1083Hooke’s law, 953hydrogen atoms-states, 1144electron wavefunction, 208ground-state energy, 800hydrogen molecule, symmetries of, 1041hyperbolaas section of quadratic surface, 292equation <strong>for</strong>, 16hyperbolic functions, 102–109, 833calculus of, 106–109definitions, 102, 833graphs, 102identities, 104in equations, 105inverses, 105graphs, 106trigonometric analogies, 102–104hyperbolic PDE, 687, 690hypergeometric distribution, 1173mean <strong>and</strong> variance, 1173hypergeometric equation, 535, 628–632as example of Sturm–Liouville equation, 566,5671317


INDEXgeneral solution, 630natural interval, 567hypergeometric functions, 628–632contiguous relations, 632integral representation, 631recurrence relations, 632special cases, 630hypothesis testing, 1277–1298errors, first <strong>and</strong> second kind, 1280generalised likelihood ratio, 1282generalised likelihood-ratio test, 1281goodness of fit, 1296Neyman–Pearson test, 1280null, 1278power, 1280rejection region, 1279simple or composite, 1278statistical tests, 1278test statistic, 1278i, j, k (unit vectors), 219i, squarerootof−1, 84identity element of a group, 1043–1046uniqueness, 1043, 1045identity matrices, 254, 255identity operator, 249images, method of, see method of imagesimaginary part or term of a complex number, 83importance sampling, in Monte Carlo methods,1012improperintegrals, 70rotations, 946–948impulses, δ-function respresentation, 441incomplete gamma function, 639independent r<strong>and</strong>om variables, 1156, 1200index of a subgroup, 1066indices, of regular singular points, 540indicial equation, 540distinct roots with non-integral difference,540–542repeated roots, 542, 546, 547roots differ by integer, 542, 546induction, proof by, 31inequalitiesamongst integrals, 72Bessel, 246, 559Schwarz, 246, 559triangle, 246, 559inertia, see also moments of inertiamoments <strong>and</strong> products, 951tensor, 951inexact differentials, 155inexact equation, 473infinite integrals, 70contour integration, 862–867infinite series, see seriesinflectiongeneral points of, 52stationary points of, 50–52inhomogeneousboundary conditions, see boundaryconditions, homogeneous <strong>and</strong>inhomogeneousdifferential equations, 490simultaneous linear equations, 293inner product in a vector space, see also scalarproductof finite dimensionality, 244<strong>and</strong> Hermitian conjugate, 258commutativity, 244distributivity over addition, 244of infinite dimensionality, 557integral equationseigenfunctions, 817eigenvalues, 808, 816Fredholm, 805from differential equations, 803homogeneous, 805linear, 804first kind, 805second kind, 805nomenclature, 804singular, 805Volterra, 805integral equations, methods <strong>for</strong>differentiation, 812Fredholm theory, 815integral trans<strong>for</strong>ms, 809–812Fourier, 809–812Laplace, 810Neumann series, 813–815Schmidt–Hilbert theory, 816–819separable (degenerate) kernels, 807integral functions, 832nintegral test <strong>for</strong> convergence of series, 128integral trans<strong>for</strong>ms, see also Fourier trans<strong>for</strong>ms<strong>and</strong> Laplace trans<strong>for</strong>msgeneral <strong>for</strong>m, 459Hankel trans<strong>for</strong>ms, 459Mellin trans<strong>for</strong>ms, 459integrals, see also integrationcomplex, see complex integralsdefinite, 59double, see multiple integralsFourier trans<strong>for</strong>m of, 444improper, 70indefinite, 62inequalities, 72, 559infinite, 70Laplace trans<strong>for</strong>m of, 456limitscontaining variables, 188fixed, 59variable, 61line, see line integralsmultiple, see multiple integralsnon-zero, 1104of vectors, 339properties, 601318


INDEXtriple, see multiple integralsundefined, 59integr<strong>and</strong>, 59integrating factor (IF), 506first-order ODE, 473–475integration, see also integralsapplications, 72–76finding the length of a curve, 73mean value of a function, 72surfaces of revolution, 74volumes of revolution, 75as area under a curve, 59as the inverse of differentiation, 61<strong>for</strong>mal definition, 59from first principles, 59in plane polar coordinates, 70logarithmic, 64multiple, see multiple integralsmultivalued functions, 865–867, 885of Fourier series, 424of functions of several real variables, seemultiple integralsof hyperbolic functions, 106–109of power series, 135of simple functions, 62of singular functions, 70of sinusoidal functions, 63, 861integration constant, 62integration, methods <strong>for</strong>by inspection, 62by parts, 67–69by substitution, 65–67t substitution, 65change of variables, see change of variablescompleting the square, 66contour, see contour integrationGaussian, 1005–1009numerical, 1000–1009partial fractions, 64reduction <strong>for</strong>mulae, 69stationary phase, 912–920steepest descents, 908–912trigonometric expansions, 63using complex numbers, 101intersection ∩, probability <strong>for</strong>, 1120, 1128intrinsic derivative, see absolute derivativeinvariant tensors, see isotropic tensorsinverse hyperbolic functions, 105inverse integral trans<strong>for</strong>msFourier, 435Laplace, 454, 884–887uniqueness, 454inverse matrices, 263–266elements, 264in solution of simultaneous linear equations,295product rule, 266properties, 265inverse of a linear operator, 249inverse of a product in a group, 1046inverse of element in a groupuniqueness, 1043, 1046inversion theorem, Fourier’s, 435inversions asimproper rotations, 946symmetry operations, 1041irregular singular points, 534irreps, 1087counting, 1095dimension n λ , 1097direct sum ⊕, 1086identity A 1 , 1100, 1104n-dimensional, 1088, 1089, 1102number in a representation, 1087, 1095one-dimensional, 1088, 1089, 1093, 1099, 1102orthogonality theorem, 1090–1092projection operators <strong>for</strong>, 1107reduction to, 1096summation rules <strong>for</strong>, 1097–1099irrotational vectors, 353isobaric ODE, 476non-linear, 521isoclines, method of, 1028, 1037isomorphic groups, 1051–1056, 1058, 1059isomorphism (mapping), 1060isotope decay, 484, 525isotropic (invariant) tensors, 944–946, 953iteration schemesconvergence of, 992–994<strong>for</strong> algebraic equations, 986–994<strong>for</strong> differential equations, 1025<strong>for</strong> integral equations, 813–816Gauss–Seidel, 996–998order of convergence, 993J ν (x), see Bessel functionsj, squarerootof−1, 84j l (x), see spherical Bessel functionsJacobiansanalogy with derivatives, 207<strong>and</strong> change of variables, 206definition intwo dimensions, 201three dimensions, 205general properties, 206in terms of a determinant, 201, 205, 207joint distributions, see bivariate distributions<strong>and</strong> multivariate distributionsJordan’s lemma, 864kernel of a homomorphism, 1060, 1063kernel of an integral trans<strong>for</strong>m, 459kernel of integral equationsdisplacement, 809Hermitian, 816of <strong>for</strong>m exp(−ixz), 810–812of linear integral equations, 804resolvent, 814, 815separable (degenerate), 807ket vector |ψ〉, 6481319


INDEXkinetic energy of oscillating system, 316Klein–Gordon equation, 711, 772Kronecker delta δ ij <strong>and</strong> orthogonality, 244Kronecker delta, δ ij , δ j i , tensor, 928, 941–946,956, 962identities, 943isotropic, 945vector products, 942Kummer function, 633kurtosis, 1150, 1227L n (x), see Laguerre polynomialsL m n (x), see associated Laguerre polynomialsL’Hôpital’s rule, 142–144Lagrange equations, 789<strong>and</strong> energy conservation, 797Lagrange undetermined multipliers, 167–173<strong>and</strong> ODE eigenvalue estimation, 792application to stationary properties of theeigenvectors of quadratic <strong>and</strong> Hermitian<strong>for</strong>ms, 290<strong>for</strong> functions of more than two variables,169–173in deriving the Boltzmann distribution,171–173integral constraints, 785with several constraints, 169–173Lagrange’s identity, 226Lagrange’s theorem, 1065<strong>and</strong> the order of a subgroup, 1062<strong>and</strong> the order of an element, 1062Lagrangian, 789, 797Laguerre equation, 535, 616–621as example of Sturm–Liouville equation, 566,619natural interval, 567, 619Laguerre polynomials L n (x), 617as special case of confluent hypergeometricfunction, 634generating function, 620graph of, 617normalisation, 619orthogonality, 619recurrence relations, 620Rodrigues’ <strong>for</strong>mula, 618Lamé constants, 953lamina: mass, centre of mass <strong>and</strong> centroid,193–195Laplace equation, 679expansion methods, 741–744in two dimensions, 688, 690, 717, 718<strong>and</strong> analytic functions, 829<strong>and</strong> con<strong>for</strong>mal trans<strong>for</strong>mations, 876–879numerical method <strong>for</strong>, 1031, 1038plane polars, 725–727separated variables, 717in three dimensionscylindrical polars, 728–731spherical polars, 731–737uniqueness of solution, 741with specified boundary values, 764, 766Laplace expansion, 259Laplace trans<strong>for</strong>ms, 453–459, 884convolutionassociativity, commutativity, distributivity,458definition, 457convolution theorem, 457definition, 453<strong>for</strong> ODE with constant coefficients, 501–503<strong>for</strong> PDE, 747–748inverse, 454, 884–887uniqueness, 454properties: translation, exponentialmultiplication, etc., 456table <strong>for</strong> common functions, 455Laplace trans<strong>for</strong>ms, examplesconstant, 453derivatives, 455exponential function, 453integrals, 456polynomial, 453Laplacian, see del squared ∇ 2 (Laplacian)Laurent expansion, 855–858analytic <strong>and</strong> principal parts, 855region of convergence, 855least squares, method of, 1271–1277basis functions, 1273linear, 1272non-linear, 1276response matrix, 1273Legendre equation, 534, 535, 577–586as example of Sturm–Liouville equation, 566,583associated, see associated Legendre equationgeneral solution, 578, 580natural interval, 567, 583Legendre functions P l (x), 577–586associated Legendre functions, 768of second kind Q l (x), 579graph of, 580Legendre linear equation, 503Legendre polynomials P l (x), 578as special case of hypergeometric function,631associated Legendre functions, 733generating function, 584–586graph of, 579in Gaussian integration, 1006normalisation, 578, 582orthogonality, 583, 735recurrence relations, 585, 586Rodrigues’ <strong>for</strong>mula, 581Leibnitz’ rule <strong>for</strong> differentiation of integrals, 178Leibnitz’ theorem, 48–50length ofa vector, 218plane curves, 73, 341space curves, 341tensor <strong>for</strong>m, 9821320


INDEXlevel lines, 905, 906Levi-Civita symbol, see ɛ ijk , Levi-Civita symbol,tensorlikelihood function, 1255limits, 141–144definition, 141L’Hôpital’s rule, 142–144of functions containing exponents, 142of integrals, 59containing variables, 188of products, 141of quotients, 141–144of sums, 141line charge, electrostatic potential, 872, 878line integrals<strong>and</strong> Cauchy integrals, 851–853<strong>and</strong> Stokes’ theorem, 406–409of scalars, 377–387of vectors, 377–389physical examples, 381round closed loop, 386line of steepest descents, 908line, vector equation of, 226linear dependence <strong>and</strong> independencedefinition in a vector space, 242of basis vectors, 217relationship with rank, 267linear differential operator L, 511, 545, 554adjoint L † , 559eigenfunctions, see eigenfunctionseigenvalues, see eigenvalues, of lineardifferential operators<strong>for</strong> Sturm-Liouville equation, 564–568Hermitian, 555, 559–564self-adjoint, 559linear equations, differentialfirst-order ODE, 474general ODE, 490–517ODE with constant coefficients, 492–503ODE with variable coefficients, 503–517linear equations, simultaneous, see simultaneouslinear equationslinear independence of functions, 491Wronskian test, 491, 532linear integral operator K, 805<strong>and</strong> Schmidt–Hilbert theory, 816–818Hermitian conjugate, 805inverse, 806linear interpolation <strong>for</strong> algebraic equations, 988linear least squares, method of, 1272linear moleculesnormal modes of, 320–322symmetries of, 1077linear operators, 247–249associativity, 249distributivity over addition, 249eigenvalues <strong>and</strong> eigenvectors, 272in a particular basis, 248inverse, 249non-commutativity, 249particular: identity, null or zero, singular <strong>and</strong>non-singular, 249properties, 249linear vector spaces, see vector spaceslines of steepest descent, 906Liouville’s theorem, 853Ln of a complex number, 99, 834ln (natural logarithm)Maclaurin series <strong>for</strong>, 140of a complex number, 99, 834log-likelihood function, 1258longitudinal vibrations in a rod, 677lottery (UK), <strong>and</strong> hypergeometric distribution,1174lower triangular matrices, 269Maclaurin series, 138st<strong>and</strong>ard expressions, 140Madelung constant, 149magnetic dipole, 220magnitude of a vector, 218in terms of scalar or dot product, 221mappings between groups, see groups, mappingsbetweenmarginal distributions, 1198mass of non-uni<strong>for</strong>m bodies, 193matrices, 241–307as a vector space, 252as arrays of numbers, 249as representation of a linear operator, 249column, 250elements, 249minors <strong>and</strong> cofactors, 259identity or unit, 254row, 250zero or null, 254matrices, algebra of, 250addition, 251change of basis, 283–285Cholesky separation, 313diagonalisation, see diagonalisation ofmatricesmultiplication, 252–254<strong>and</strong> common eigenvalues, 278commutator, 309non-commutativity, 254multiplication by a scalar, 251normal modes, see normal modesnumerical methods, see numerical methods<strong>for</strong> simultaneous linear equationssimilarity trans<strong>for</strong>mations, see similaritytrans<strong>for</strong>mationssimultaneous linear equations, seesimultaneous linear equationssubtraction, 251matrices, derivedadjoint, 256–258complex conjugate, 256–258Hermitian conjugate, 256–258inverse, see inverse matrices1321


INDEXtranspose, 250matrices, properties ofanti- or skew-symmetric, 270anti-Hermitian, see anti-Hermitian matricesdeterminant, see determinantsdiagonal, 268eigenvalues, see eigenvalueseigenvectors, see eigenvectorsHermitian, see Hermitian matricesnormal, see normal matricesnullity, 293order, 249orthogonal, 270rank, 267square, 249symmetric, 270trace or spur, 258triangular, 269tridiagonal, 998–1000, 1030unitary, see unitary matricesmatrix elements in quantum mechanicsas integrals, 1103dipole, 1108, 1115maxima <strong>and</strong> minima (local) of a function ofconstrained variables, see Lagrangeundetermined multipliersone real variable, 50–52sufficient conditions, 51several real variables, 162–167sufficient conditions, 164, 167maximum modulus theorem, 881maximum-likelihood, method of, 1255–1271<strong>and</strong> Bayesian approach, 1264bias, 1260data modelling, 1255estimator, 1256extended, 1270log-likelihood function, 1258parameter estimation, 1255trans<strong>for</strong>mation invariance, 1260Maxwell’selectromagnetic equations, 373, 408, 979thermodynamic relations, 176–178Maxwell–Boltzmann statistics, 1138mean µfrom MGF, 1163from PGF, 1158of RVD, 1144of sample, 1223of sample: geometric, harmonic, root meansquare, 1223mean value of a function ofone variable, 72several variables, 199mean value theorem, 56median of RVD, 1145membranede<strong>for</strong>med rim, 725–727normal modes, 739, 1112transverse vibrations, 677, 739, 768, 1112method of images, 706, 758–765, 878disc (section of cylinder), 764, 766infinite plate, 759intersecting plates in two dimensions, 761sphere, 762–764, 772metric tensor, 957–960, 963<strong>and</strong> Christoffel symbols, 966<strong>and</strong> scale factors, 957, 972covariant derivative of, 982determinant, 957, 964derivative of, 973length element, 957raising or lowering index, 959, 963scalar product, 958volume element, 957, 981MGF, see moment generating functionsMilne’s method, 1022minimum-variance estimator, 1232minor of a matrix element, 259mixed, components of tensor, 957, 962, 969ML estimators, 1256bias, 1260confidence limits, 1262efficiency, 1261trans<strong>for</strong>mation invariance, 1260mode of RVD, 1145modulo, mod N, multiplication, 1049modulusof a complex number, 87of a vector, see magnitude of a vectormoleculesbonding in, 1103, 1105–1108dipole moments of, 1077symmetries of, 1077moment generating functions (MGFs),1162–1167<strong>and</strong> central limit theorem, 1195<strong>and</strong> PGF, 1163mean <strong>and</strong> variance, 1163particular distributionsbinomial, 1170exponential, 1191Gaussian, 1163, 1185Poisson, 1177properties, 1163moments (of distributions)central, 1148of RVD, 1147moments (of <strong>for</strong>ces), vector representation of,223moments of inertia<strong>and</strong> inertia tensor, 951definition, 198of disc, 208of rectangular lamina, 198of right circular cylinder, 209of sphere, 205perpendicular axes theorem, 209momentum as first-order tensor, 933monomorphism, 10611322


INDEXMonte Carlo methods, of integration, 1009–1017antithetic variates, 1014control variates, 1013crude, 1011hit or miss, 1014importance sampling, 1012multiple integrals, 1016r<strong>and</strong>om number generation, 1017stratified sampling, 1012Morera’s theorem, 851multinomial distribution, 1208<strong>and</strong> multiple Poisson distribution, 1218multiple angles, trigonometric <strong>for</strong>mulae, 10multiple integralsapplication in findingarea <strong>and</strong> volume, 191–193mass, centre of mass <strong>and</strong> centroid, 193–195mean value of a function of severalvariables, 199moments of inertia, 198change of variablesdouble integrals, 200–204general properties, 206triple integrals, 204definitions ofdouble integrals, 187triple integrals, 190evaluation, 188–190notation, 188, 189, 191order of integration, 188, 191multiplication tables <strong>for</strong> groups, see groupmultiplication tablesmultiplication theorem, see Parseval’s theoremmultivalued functions, 835–837integration of, 865–867multivariate distributions, 1196, 1207–1211change of variables, 1206Gaussian, 1209multinomial, 1208mutually exclusive events, 1120, 1129n l (x), see spherical Bessel functionsnabla ∇, see gradient operator (grad)natural interval<strong>for</strong> associated Laguerre equation, 567, 622<strong>for</strong> associated Legendre equation, 567, 590,591<strong>for</strong> Bessel equation, 608<strong>for</strong> Chebyshev equation, 567, 599<strong>for</strong> Hermite equation, 567<strong>for</strong> Laguerre equation, 567, 619<strong>for</strong> Legendre equation, 567, 583<strong>for</strong> simple harmonic oscillator equation, 567<strong>for</strong> Sturm–Liouville equations, 565, 567natural logarithm, see ln <strong>and</strong> Lnnatural numbers, in series, 31, 121natural representations, 1081, 1110necessary <strong>and</strong> sufficient conditions, 34negative binomial distribution, 1172negative function, 556negative vector, 242Neumann boundary conditions, 702Green’s functions, 754, 765–767method of images, 765–767self-consistency, 765Neumann functions Y ν (x), 607Neumann series, 813–815Newton–Raphson (NR) method, 990–992order of convergence, 993Neyman–Pearson test, 1280nodes of oscillation, 693non-Abelian groups, 1052–1056of functions, 1055of matrices, 1054of permutations, 1056–1058of rotations <strong>and</strong> reflections, 1052non-Cartesian coordinates, see curvilinear,cylindrical polar, plane polar <strong>and</strong> sphericalpolar coordinatesnon-linear differential equations, see ordinarydifferential equations, non-linearnon-linear least squares, method of, 1276norm offunction, 557vector, 244normalto coordinate surface, 366to plane, 228to surface, 346, 350, 390normal derivative, 350normal distribution, see Gaussian (normal)distributionnormal matrices, 272eigenvectorscompleteness, 275orthogonality, 275eigenvectors <strong>and</strong> eigenvalues, 273–276normal modes, 316–329characteristic equation, 319coupled pendulums, 329, 331definition, 320degeneracy, 1110–1113frequencies of, 319linear molecular system, 320–322membrane, 739, 1112normal coordinates, 320normal equations, 320rod–string system, 317–320symmetries of, 322normal subgroups, 1063normalisation ofeigenfunctions, 562eigenvectors, 273functions, 557vectors, 219null (zero)matrix, 254, 255operator, 249space, of a matrix, 293vector, 214, 242, 5561323


INDEXnull operation, as identity element of group,1044nullity, of a matrix, 293numerical methods <strong>for</strong> algebraic equations,985–992binary chopping, 990convergence of iteration schemes, 992–994linear interpolation, 988Newton–Raphson, 990–992rearrangement methods, 987numerical methods <strong>for</strong> integration, 1000–1009Gaussian integration, 1005–1009mid-point rule, 1034Monte Carlo, 1009nomenclature, 1001Simpson’s rule, 1004trapezium rule, 1002–1004numerical methods <strong>for</strong> ordinary differentialequations, 1020–1030accuracy <strong>and</strong> convergence, 1021Adams method, 1024difference schemes, 1021–1023Euler method, 1021first-order equations, 1021–1028higher-order equations, 1028–1030isoclines, 1028Milne’s method, 1022prediction <strong>and</strong> correction, 1024–1026reduction to matrix <strong>for</strong>m, 1030Runge–Kutta methods, 1026–1028Taylor series methods, 1023numerical methods <strong>for</strong> partial differentialequations, 1030–1032diffusion equation, 1032Laplace’s equation, 1031minimising error, 1032numerical methods <strong>for</strong> simultaneous linearequations, 994–1000Gauss–Seidel iteration, 996–998Gaussian elimination with interchange, 995matrix <strong>for</strong>m, 994–1000tridiagonal matrices, 998–1000O(x), order of, 132observables in quantum mechanics, 277, 560odd functions, see antisymmetric functionsODE, see ordinary differential equations (ODEs)operatorsHermitian, see Hermitian operatorslinear, see linear operators <strong>and</strong> lineardifferential operator <strong>and</strong> linear integraloperatoroperators (quantum)angular momentum, 656–663annihilation <strong>and</strong> creation, 667coordinate-free, 648–671eigenvalues <strong>and</strong> eigenstates, 649physical examplesangular momentum, 658Hamiltonian, 657order ofapproximation in Taylor series, 137nconvergence of iteration schemes, 993group, 1043group element, 1047ODE, 468permutation, 1058recurrence relations (series), 497subgroup, 1061<strong>and</strong> Lagrange’s theorem, 1065tensor, 930ordinary differential equations (ODE), see alsodifferential equations, particularboundary conditions, 468, 470, 501complementary function, 491degree, 468dimensionally homogeneous, 475exact, 472, 505first-order, 468–484first-order higher-degree, 480–484soluble <strong>for</strong> p, 480soluble <strong>for</strong> x, 481soluble <strong>for</strong> y, 482general <strong>for</strong>m of solution, 468–470higher-order, 490–523homogeneous, 490inexact, 473isobaric, 476, 521linear, 474, 490–517non-linear, 518–523exact, 519isobaric (homogeneous), 521x absent, 518y absent, 518order, 468ordinary point, see ordinary points of ODEparticular integral (solution), 469, 492, 494singular point, see singular points of ODEsingular solution, 469, 481, 482, 484ordinary differential equations, methods <strong>for</strong>canonical <strong>for</strong>m <strong>for</strong> second-order equations,516eigenfunctions, 554–573equations containing linear <strong>for</strong>ms, 478–480equations with constant coefficients, 492–503Green’s functions, 511–516integrating factors, 473–475Laplace trans<strong>for</strong>ms, 501–503numerical, 1020–1030partially known CF, 506separable variables, 471series solutions, 531–550, 604undetermined coefficients, 494variation of parameters, 508–510ordinary points of ODE, 533, 535–538indicial equation, 543orthogonal lines, condition <strong>for</strong>, 12orthogonal matrices, 270, 929, 930general properties, see unitary matricesorthogonal systems of coordinates, 3641324


INDEXorthogonal trans<strong>for</strong>mations, 932orthogonalisation (Gram–Schmidt) ofeigenfunctions of an Hermitian operator, 562eigenvectors of a normal matrix, 275functions in a Hilbert space, 557orthogonality ofeigenfunctions of an Hermitian operator,561–563eigenvectors of a normal matrix, 275eigenvectors of an Hermitian matrix, 277functions, 557terms in Fourier series, 417, 425vectors, 219, 244orthogonality properties of characters, 1094,1102orthogonality theorem <strong>for</strong> irreps, 1090–1092orthonormalbasis functions, 557basis vectors, 244under unitary trans<strong>for</strong>mation, 285oscillations, see normal modesoutcome, of trial, 1119outer product of two vectors, 936P l (x), see Legendre polynomialsPl m (x), see associated Legendre functionsPappus’ theorems, 195–197parabola, equation <strong>for</strong>, 16parabolic PDE, 687, 690parallel axis theorem, 238parallel vectors, 223parallelepiped, volume of, 225parallelogram equality, 247parallelogram, area of, 223, 224parameter estimation (statistics), 1229–1255,1298Bessel correction, 1248error in mean, 1298maximum-likelihood, 1255mean, 1243variance, 1245–1248parameters, variation of, 508–510parametric equationsof conic sections, 17of cycloid, 370, 785of space curves, 340of surfaces, 345parity inversion, 1102Parseval’s theoremconservation of energy, 451<strong>for</strong> Fourier series, 426<strong>for</strong> Fourier trans<strong>for</strong>ms, 450partial derivative, see partial differentiationpartial differential equations (PDE), 675–707,713–767, see also differential equations,particulararbitrary functions, 680–685boundary conditions, 681, 699–707, 723characteristics, 699–705<strong>and</strong> equation type, 703equation types, 687, 710first-order, 681–687general solution, 681–692homogeneous, 685inhomogeneous equation <strong>and</strong> problem,685–687, 744–746, 751–767particular solutions (integrals), 685–692second-order, 687–698partial differential equations (PDE), methods <strong>for</strong>change of variables, 691, 696–698constant coefficients, 687general solution, 689integral trans<strong>for</strong>m methods, 747–751method of images, see method of imagesnumerical, 1030–1032separation of variables, see separation ofvariablessuperposition methods, 717–724with no undifferentiated term, 684partial differentiation, 151–179as gradient of a function of several realvariables, 151chain rule, 157change of variables, 158–160definitions, 151–153properties, 157cyclic relation, 157reciprocity relation, 157partial fractions, 18–25<strong>and</strong> degree of numerator, 21as a means of integration, 64complex roots, 22in inverse Laplace trans<strong>for</strong>ms, 454, 502repeated roots, 23partial sum, 115particular integrals (PI), 469, see also ordinarydifferential equation, methods <strong>for</strong> <strong>and</strong>partial differential equations, methods <strong>for</strong>partition of agroup, 1064set, 1065parts, integration by, 67–69path integrals, see line integralsPDE, see partial differential equationsPDFs, 1140pendulums, coupled, 329, 331periodic function representation, see Fourierseriespermutation groups S n , 1056–1058cycle notation, 1057permutation law in a group, 1047permutations, 1133–1139degree, 1056distinguishable, 1135order of, 1058symbol n P k , 1133perpendicular axes theorem, 209perpendicular vectors, 219, 244PF, see probability functionsPGF, see probability generating functions1325


INDEXphase memory, 895phase, complex, 896PI, see particular integralsplane curves, length of, 73in Cartesian coordinates, 73in plane polar coordinates, 74plane polar coordinates, 70, 336arc length, 74, 361area element, 202, 361basis vectors, 336velocity <strong>and</strong> acceleration, 337plane waves, 695, 716planes<strong>and</strong> simultaneous linear equations, 300vector equation of, 227plates, conducting, see also complex potentials,<strong>for</strong> platesline charge near, 761point charge near, 759point charges, δ-function respresentation, 441point groups, 1082points of inflection of a function ofone real variable, 50–52several real variables, 162–167Poisson distribution Po(λ), 1174–1179<strong>and</strong> Gaussian distribution, 1187as limit of binomial distribution, 1174, 1177mean <strong>and</strong> variance, 1176MGF, 1177multiple, 1178recurrence <strong>for</strong>mula, 1176Poisson equation, 575, 679, 744–746fundamental solution, 757Green’s functions, 753–767uniqueness, 705–707Poisson summation <strong>for</strong>mula, 461Poisson’s ratio, 953polar coordinates, see plane polar <strong>and</strong> cylindricalpolar <strong>and</strong> spherical polar coordinatespolar representation of complex numbers, 92–95polar vectors, 949pole, of a function of a complex variablecontours containing, 861–867, 884–887order, 837, 856residue, 856–858polynomial equations, 1–10conjugate roots, 99factorisation, 7multiplicities of roots, 4number of roots, 83, 85, 868properties of roots, 9real roots, 1solution of, using de Moivre’s theorem, 98polynomial solutions of ODE, 538, 548–550populations, sampling of, 1222positive (semi-) definite quadratic <strong>and</strong> Hermitian<strong>for</strong>ms, 290positive semi-definite norm, 244potential energy ofion in a crystal lattice, 148magnetic dipole in a field, 220oscillating system, 317potential function<strong>and</strong> conservative fields, 389complex, 871–876electrostatic, see electrostatic fields <strong>and</strong>potentialsgravitational, see gravitational fields <strong>and</strong>potentialsvector, 389power series<strong>and</strong> differential equations, see series solutionsof differential equationsinterval of convergence, 132Maclaurin, see Maclaurin seriesmanipulation: difference, differentiation,integration, product, substitution, sum, 134Taylor, see Taylor seriespower series in a complex variable, 133, 830–832analyticity, 832circle <strong>and</strong> radius of convergence, 133, 831convergence tests, 831, 832<strong>for</strong>m, 830power, in hypothesis testing, 1280powers, complex, 99, 833prediction <strong>and</strong> correction methods, 1024–1026,1035prime, non-existence of largest, 34principal axes ofCartesian tensors, 951–953conductivity tensors, 952inertia tensors, 951quadratic surfaces, 292rotation symmetry, 1102principal normals of space curves, 342principal value ofcomplex integrals, 864complex logarithms, 100, 834principle of the argument, 880probability, 1124–1211axioms, 1125conditional, 1128–1133Bayes’ theorem, 1132combining, 1130definition, 1125<strong>for</strong> intersection ∩, 1120<strong>for</strong> union ∪, 1121, 1125–1128probability distributions, 1139, see alsoindividual distributionsbivariate, see bivariate distributionschange of variables, 1150–1157cumulants, 1166generating functions, see moment generatingfunctions <strong>and</strong> probability generatingfunctionsmean µ, 1144mean of functions, 1145mode, median <strong>and</strong> quartiles, 1145moments, 1147–1150multivariate, see multivariate distributions1326


INDEXst<strong>and</strong>ard deviation σ, 1146variance σ 2 , 1146probability functions (PFs), 1139cumulative (CPFs), 1139, 1141density functions (PDFs), 1140probability generating functions (PGFs),1157–1162<strong>and</strong> MGF, 1163binomial, 1161definition, 1158geometric, 1159mean <strong>and</strong> variance, 1158Poisson, 1158sums of RV, 1161trials, 1158variable sums of RV, 1161product rule <strong>for</strong> differentiation, 44–46, 48–50products of inertia, 951projection operators <strong>for</strong> irreps, 1107, 1116projection tensors, 979proper rotations, 946proper subgroups, 1061pseudoscalars, 947, 950pseudotensors, 946–950, 964pseudovectors, 946–950Q l (x), see Legendre polynomialsQ m l (x), see associated Legendre functionsquadratic equationscomplex roots of, 83properties of roots, 10roots of, 2quadratic <strong>for</strong>ms, 288–292completing the square, 1206positive definite <strong>and</strong> semi-definite, 290quadratic surfaces, 292removing cross terms, 289stationary properties of eigenvectors, 290quantum mechanics, from classical mechanics,657quantum operators, see operators (quantum)quartiles, of RVD, 1145quaternion group, 1073, 1113quotient law <strong>for</strong> tensors, 939–941quotient rule <strong>for</strong> differentiation, 47quotient test <strong>for</strong> series, 127radius of convergence, 133, 831radius of curvatureof plane curves, 53of space curves, 342radius of torsion of space curves, 343r<strong>and</strong>om number generation, 1017r<strong>and</strong>om numbers, non-uni<strong>for</strong>m distribution,1035r<strong>and</strong>om variable distributions, see probabilitydistributionsr<strong>and</strong>om variables (RV), 1119, 1139–1143continuous, 1140–1143dependent, 1196–1205discrete, 1139independent, 1156, 1200sums of, 1160–1162uncorrelated, 1200range of a matrix, 293rank of matrices, 267<strong>and</strong> determinants, 267<strong>and</strong> linear dependence, 267rank of tensors, see order of tensorrate of change of a function ofone real variable, 41several real variables, 153–155ratio comparison test, 127ratio test (D’Alembert), 126, 832in convergence of power series, 132ratio theorem, 215<strong>and</strong> centroid of a triangle, 216Rayleigh–Ritz method, 327–329, 800real part or term of a complex number, 83real roots, of a polynomial equation, 1rearrangement methods <strong>for</strong> algebraic equations,987reciprocal vectors, 233, 366, 955, 959reciprocity relation <strong>for</strong> partial derivatives, 157rectangular distribution, 1194Fourier trans<strong>for</strong>m of, 442recurrence relations (functions), 585, 611associated Laguerre polynomials, 624associated Legendre functions, 592Chebyshev polynomials, 601confluent hypergeometric functions, 635Hermite polynomials, 628hypergeometric functions, 632Laguerre polynomials, 620Legendre polynomials, 586recurrence relations (series), 496–501characteristic equation, 499coefficients, 536, 538, 999first-order, 497second-order, 499higher-order, 501reducible representations, 1084, 1086reduction <strong>for</strong>mulae <strong>for</strong> integrals, 69reflections<strong>and</strong> improper rotations, 946as symmetry operations, 1041reflexivity, <strong>and</strong> equivalence relations, 1064regular functions, see analytic functionsregular representations, 1097, 1110regular singular points, 534, 538–540relative velocities, 218remainder term in Taylor series, 138repeated roots of auxiliary equation, 493representation, 1076definition, 1082dimension of, 1078, 1082equivalent, 1084–1086faithful, 1083, 1098generation of, 1078–1084, 1112irreducible, see irreps1327


INDEXnatural, 1081, 1110product, 1103–1105reducible, 1084, 1086regular, 1097, 1110counting irreps, 1098unitary, 1086representative matrices, 1079block-diagonal, 1086eigenvalues, 1100inverse, 1083number needed, <strong>and</strong> order of group, 1082of identity, 1082residueat a pole, 856–858theorem, 858–860resolution function, 446resolvent kernel, 814, 815response matrix, <strong>for</strong> linear least squares, 1273rhomboid, volume of, 237Riemann tensor, 981Riemann theorem <strong>for</strong> conditional convergence,124Riemann zeta series, 128, 129right h<strong>and</strong> screw rule, 222Rodrigues’ <strong>for</strong>mula <strong>for</strong>associated Laguerre polynomials, 622associated Legendre functions, 588Chebyshev polynomials, 599Hermite polynomials, 626Laguerre polynomials, 618Legendre polynomials, 581Rolle’s theorem, 55root test (Cauchy), 129, 831rootsof a polynomial equation, 2properties, 9of unity, 97rope, suspended at its ends, 786rotation groups (continuous), invariantsubspaces, 1088rotation matrices as a group, 1048rotation of a vector, see curlrotationsas symmetry operations, 1041axes <strong>and</strong> orthogonal matrices, 930, 931, 961improper, 946–948invariance under, 934product of, 931proper, 946Rouché’s theorem, 880–882row matrix, 250Runge–Kutta methods, 1026–1028RV, see r<strong>and</strong>om variablesRVD (r<strong>and</strong>om variable distributions), seeprobability distributionssaddle point method of integration, 908saddle points, 162<strong>and</strong> integral evaluation, 905sufficient conditions, 164, 167samplingcorrelation, 1227covariance, 1227space, 1119statistics, 1222–1229with or without replacement, 1129scalar fields, 347derivative along a space curve, 349gradient, 348–352line integrals, 377–387rate of change, 349scalar product, 219–222<strong>and</strong> inner product, 244<strong>and</strong> metric tensor, 958<strong>and</strong> perpendicular vectors, 219, 244<strong>for</strong> vectors with complex components, 221in Cartesian coordinates, 221invariance, 930, 939scalar triple product, 224–226cyclic permutation of, 225in Cartesian coordinates, 225determinant <strong>for</strong>m, 225interchange of dot <strong>and</strong> cross, 225scalars, 212invariance, 930zero-order tensors, 933scale factors, 359, 362, 364<strong>and</strong> metric tensor, 957, 972scattering in quantum mechanics, 463Schmidt–Hilbert theory, 816–819Schrödinger equation, 679constant potential, 768hydrogen atom, 741numerical solution, 1039variational approach, 795Schwarz inequality, 246, 559Schwarz–Christoffel trans<strong>for</strong>mation, 843second differences, 1019second-order differential equations, see ordinarydifferential equations <strong>and</strong> partial differentialequationssecular determinant, 280self-adjoint operators, 559–564, see alsoHermitian operatorssemicircle, angle in, 18semicircular lamina, centre of mass, 197separable kernel in integral equations, 807separable variables in ODE, 471separation constants, 715, 717separation of variables, <strong>for</strong> PDE, 713–746diffusion equation, 716, 722–724, 737, 751expansion methods, 741–744general method, 713–717Helmholtz equation, 737–741inhomogeneous boundary conditions, 722–724inhomogeneous equations, 744–746Laplace equation, 717–722, 725–737, 741polar coordinates, 725–746separation constants, 715, 717superposition methods, 717–7241328


INDEXwave equation, 714–716, 737, 739series, 115–141convergence of, see convergence of infiniteseriesdifferentiation of, 131finite <strong>and</strong> infinite, 116integration of, 131multiplication by a scalar, 131multiplication of (Cauchy product), 131notation, 116operations, 131summation, see summation of seriesseries, particulararithmetic, 117arithmetico-geometric, 118Fourier, see Fourier seriesgeometric, 117Maclaurin, 138, 140power, see power seriespowers of natural numbers, 121Riemann zeta, 128, 129Taylor, see Taylor seriesseries solutions of differential equations, 531–550about ordinary points, 535–538about regular singular points, 538–540Frobenius series, 539convergence, 536indicial equation, 540linear independence, 540polynomial solutions, 538, 548–550recurrence relation, 536, 538second solution, 537, 544–548derivative method, 545–548Wronskian method, 544, 580shortest path, 778<strong>and</strong> geodesics, 976, 982similarity trans<strong>for</strong>mations, 283–285, 929, 1092properties of matrix under, 284unitary trans<strong>for</strong>mations, 285simple harmonic oscillator, 555energy levels of, 642, 902equation, 535, 566operator <strong>for</strong>malism, 667simple poles, 838Simpson’s rule, 1004simultaneous linear equations, 292–307<strong>and</strong> intersection of planes, 300homogeneous <strong>and</strong> inhomogeneous, 293singular value decomposition, 301–307solution usingCramer’s rule, 299inverse matrix, 295numerical methods, see numerical methods<strong>for</strong> simultaneous linear equationssine, sin(x)in terms of exponential functions, 102Maclaurin series <strong>for</strong>, 140orthogonality relations, 417singular <strong>and</strong> non-singularintegral equations, 805linear operators, 249matrices, 263singular integrals, see improper integralssingular point (singularity), 826, 837–839essential, 838, 856removable, 838singular points of ODE, 533irregular, 534particular equations, 535regular, 534singular solution of ODE, 469, 481, 482, 484singular value decomposition<strong>and</strong> simultaneous linear equations, 301–307singular values, 302sinh, hyperbolic sine, 102, 833, see alsohyperbolic functionsskew-symmetric matrices, 270skewness, 1150, 1227Snell’s law, 788soap films, 780solenoidal vectors, 352, 389solid angleas surface integral, 395subtended by rectangle, 411solid: mass, centre of mass <strong>and</strong> centroid,193–195source density, 679space curves, 340–344arc length, 341binormal, 342curvature, 342Frenet–Serret <strong>for</strong>mulae, 343parametric equations, 340principal normal, 342radius of curvature, 342radius of torsion, 343tangent vector, 342torsion, 342spaces, see vector spacesspan of a set of vectors, 242sphere, vector equation of, 228spherical Bessel functions, 615, 741of first kind j l (z), 615of second kind n l (z), 615spherical harmonics Yl m (θ, φ), 593–595addition theorem, 594spherical polar coordinates, 361–363area element, 362basis vectors, 362length element, 362vector operators, 361–363volume element, 205, 362spur of a matrix, see trace of a matrixsquare matrices, 249square, symmetries of, 1100square-wave, Fourier series <strong>for</strong>, 418stagnation points of fluid flow, 873st<strong>and</strong>ard deviation σ, 1146of sample, 1224st<strong>and</strong>ing waves, 6931329


INDEXstationary phase, method of, 912–920stationary valuesof functions ofone real variable, 50–52several real variables, 162–167of integrals, 776under constraints, see Lagrange undeterminedmultipliersstatistical tests, <strong>and</strong> hypothesis testing , 1278statistics, 1119, 1221–1298describing data, 1222–1229estimating parameters, 1229–1255, 1298steepest descents, method of, 908–912Stirling’sapproximation, 637, 1185asymptotic series, 637Stokes constant, in Stokes phenomenon, 904Stokes line, 899, 903Stokes phenomenon, 903dominant term, 904Stokes constant, 904subdominant term, 904Stokes’ equation, 643, 799, 888–894Airy integrals, 890–894qualitative solutions, 888series solution, 890Stokes’ theorem, 388, 406–409<strong>for</strong> tensors, 955physical applications, 408related theorems, 407strain tensor, 953stratified sampling, in Monte Carlo methods,1012streamlines <strong>and</strong> complex potentials, 873stress tensor, 953stress waves, 980stringloaded, 798plucked, 770transverse vibrations of, 676, 789Student’s t-distributionnormalisation, 1286plots, 1287Student’s t-test, 1284–1290comparison of means, 1289critical points table, 1288one- <strong>and</strong> two-tailed confidence limits, 1288Sturm–Liouville equations, 564boundary conditions, 564examples, 566associated Laguerre, 566, 622associated Legendre, 566, 590, 591Bessel, 566, 608–611Chebyshev, 566, 599confluent hypergeometric, 566Hermite, 566hypergeometric, 566Laguerre, 566, 619Legendre, 566, 583manipulation to self-adjoint <strong>for</strong>m, 565–568natural interval, 565, 567two independent variables, 801variational approach, 790–795weight function, 790zeros of eigenfunctions, 573subdominant term, in Stokes phenomenon, 904subgroups, 1061–1063index, 1066normal, 1063order, 1061Lagrange’s theorem, 1065proper, 1061trivial, 1061submatrices, 267subscripts <strong>and</strong> superscripts, 928contra- <strong>and</strong> covariant, 956covariant derivative, 969dummy, 928free, 928partial derivative, 969summation convention, 928, 955substitution, integration by, 65–67summation convention, 928, 955summation of series, 116–124arithmetic, 117arithmetico-geometric, 118contour integration method, 882difference method, 119Fourier series method, 427geometric, 117powers of natural numbers, 121trans<strong>for</strong>mation methods, 122–124differentiation, 122integration, 122substitution, 123superposition methods<strong>for</strong> ODE, 554, 568–571<strong>for</strong> PDE, 717–724surface integrals<strong>and</strong> divergence theorem, 401Archimedean upthrust, 396, 410of scalars, vectors, 389–396physical examples, 395surfaces, 345–347area of, 346cone, 74solid, <strong>and</strong> Pappus’ theorem, 195–197sphere, 346coordinate curves, 346normal to, 346, 350of revolution, 74parametric equations, 345quadratic, 292tangent plane, 346SVD, see singular value decompositionsymmetric functions, 416<strong>and</strong> Fourier series, 419<strong>and</strong> Fourier trans<strong>for</strong>ms, 445symmetric matrices, 270general properties, see Hermitian matrices1330


INDEXsymmetric tensors, 938symmetry in equivalence relations, 1064symmetry operationson molecules, 1041order of application, 1044t-test, see Student’s t-testt substitution, 65tan −1 x, Maclaurin series <strong>for</strong>, 140tangent planes to surfaces, 346tangent vectors to space curves, 342tanh, hyperbolic tangent, see hyperbolicfunctionsTaylor series, 136–141<strong>and</strong> finite differences, 1019, 1026<strong>and</strong> Taylor’s theorem, 136–139, 853approximation errors, 139in numerical methods, 992, 1003as solution of ODE, 1023<strong>for</strong> functions of a complex variable, 853–855<strong>for</strong> functions of several real variables, 160–162remainder term, 138required properties, 136st<strong>and</strong>ard <strong>for</strong>ms, 136tensors, see Cartesian tensors <strong>and</strong> Cartesiantensors, particular <strong>and</strong> general tensorstest statistic, 1278tetrahedral group, 1115tetrahedronmass of, 194volume of, 192thermodynamicsfirst law of, 176Maxwell’s relations, 176–178top-hat function, see rectangular distributiontorque, vector representation of, 223torsion of space curves, 342total derivative, 154total differential, 154trace of a matrix, 258<strong>and</strong> second-order tensors, 939as sum of eigenvalues, 280, 287invariance under similarity trans<strong>for</strong>mations,284, 1092trace <strong>for</strong>mula, 287transcendental equations, 986trans<strong>for</strong>mation matrix, 283, 289trans<strong>for</strong>mationsactive <strong>and</strong> passive, 948con<strong>for</strong>mal, 839–844coordinate, see coordinate trans<strong>for</strong>mationssimilarity, see similarity trans<strong>for</strong>mationstrans<strong>for</strong>ms, integral, see integral trans<strong>for</strong>ms <strong>and</strong>Fourier trans<strong>for</strong>ms <strong>and</strong> Laplace trans<strong>for</strong>mstransients, <strong>and</strong> the diffusion equation, 723transitivity in equivalence relations, 1064transpose of a matrix, 250, 255product rule, 256transverse vibrationsmembrane, 677, 739, 768rod, 769string, 676trapezium rule, 1002–1004trial functions<strong>for</strong> eigenvalue estimation, 793<strong>for</strong> particular integrals of ODE, 494trials, 1119triangle inequality, 246, 559triangle, centroid of, 216triangular matrices, 269tridiagonal matrices, 998–1000, 1030, 1033trigonometric identities, 10–15triple integrals, see multiple integralstriple scalar product, see scalar triple producttriple vector product, see vector triple productturning point, 50uncertainty principle (Heisenberg), 435–437from commutator, 664undetermined coefficients, method of, 494undetermined multipliers, see Lagrangeundetermined multipliersuni<strong>for</strong>m distribution, 1194union ∪, probability <strong>for</strong>, 1121, 1125, 1128uniqueness theoremKlein–Gordon equation, 711Laplace equation, 741Poisson equation, 705–707unit step function, see Heaviside functionunit vectors, 219unitarymatrices, 271eigenvalues <strong>and</strong> eigenvectors, 278representations, 1086trans<strong>for</strong>mations, 285upper triangular matrices, 269variable end-points, 782–785variable, dummy, 61variables, separation of, see separation ofvariablesvariance σ 2 , 1146from MGF, 1163from PGF, 1159of dependent RV, 1203of sample, 1224variation of parameters, 508–510variation, constrained, 785–787variational principles, physical, 787–790Fermat, 787Hamilton, 788variations, calculus of, see calculus of variationsvector operators, 347–369acting on sums <strong>and</strong> products, 354combinations of, 355–357curl, 353, 368del ∇, 348del squared ∇ 2 , 352divergence (div), 352geometrical definitions, 398–4001331


INDEXgradient operator (grad), 348–352, 367identities, 356, 978Laplacian, 352, 368non-Cartesian, 357–369tensor <strong>for</strong>ms, 971–975curl, 974divergence, 972gradient, 972Laplacian, 973vector product, 222–224anticommutativity, 222definition, 222determinant <strong>for</strong>m, 224in Cartesian coordinates, 224non-associativity, 222vector spaces, 242–247, 1113associativity of addition, 242basis vectors, 243commutativity of addition, 242complex, 242defining properties, 242dimensionality, 243group actions on, 1088inequalities: Bessel, Schwarz, triangle, 246invariant, 1088, 1113matrices as an example, 252of infinite dimensionality, 556–559associativity of addition, 556basis functions, 556commutativity of addition, 556defining properties, 556Hilbert spaces, 557–559inequalities: Bessel, Schwarz, triangle, 559parallelogram equality, 247real, 242span of a set of vectors in, 242vector triple product, 226identities, 226non-associativity, 226vectorsas first-order tensors, 932as geometrical objects, 241base, 336column, 250compared with scalars, 212component <strong>for</strong>m, 217examples of, 212graphical representation of, 212irrotational, 353magnitude of, 218non-Cartesian, 336, 358, 362notation, 212polar <strong>and</strong> axial, 949solenoidal, 352, 389span of, 242vectors, algebra of, 212–234addition <strong>and</strong> subtraction, 213in component <strong>for</strong>m, 218angle between, 221associativity of addition <strong>and</strong> subtraction, 213commutativity of addition <strong>and</strong> subtraction,213multiplication by a complex scalar, 222multiplication by a scalar, 214multiplication of, see scalar product <strong>and</strong>vector productouter product, 936vectors, applicationscentroid of a triangle, 216equation of a line, 226equation of a plane, 227equation of a sphere, 228finding distance from aline to a line, 231line to a plane, 232point to a line, 229point to a plane, 230intersection of two planes, 228vectors, calculus of, 334–369differentiation, 334–339, 344integration, 339line integrals, 377–389surface integrals, 389–396volume integrals, 396vectors, derived quantitiescurl, 353derivative, 334differential, 338, 344divergence (div), 352reciprocal, 233, 366, 955, 959vector fields, 347curl, 406divergence, 352flux, 395rate of change, 350vectors, physicalacceleration, 335angular momentum, 238angular velocity, 223, 238, 353area, 393–395, 408area of parallelogram, 223, 224<strong>for</strong>ce, 212, 213, 220moment or torque of a <strong>for</strong>ce, 223velocity, 335velocity vectors, 335Venn diagrams, 1119–1124vibrationsinternal, see normal modeslongitudinal, in a rod, 677transversemembrane, 677, 739, 768, 799, 801rod, 769string, 676, 789Volterra integral equation, 804, 805differentiation methods, 812Laplace trans<strong>for</strong>m methods, 810volume elementscurvilinear coordinates, 365cylindrical polars, 360spherical polars, 205, 3621332


INDEXvolume integrals, 396<strong>and</strong> divergence theorem, 401volume ofcone, 75ellipsoid, 207parallelepiped, 225rhomboid, 237tetrahedron, 192volumesas surface integrals, 397, 401in many dimensions, 210of regions, using multiple integrals, 191–193volumes of revolution, 75<strong>and</strong> surface area & centroid, 195–197zero-order tensors, 932–935zeros of a function of a complex variable, 839location of, 879–882, 921order, 839, 856principle of the argument, 880Rouché’s theorem, 880, 882zeros of Sturm-Liouville eigenfunctions, 573zeros, of a polynomial, 2zeta series (Riemann), 128, 129z-plane, see Arg<strong>and</strong> diagramwave equation, 676, 688, 790boundary conditions, 693–695characteristics, 704from Maxwell’s equations, 373in one dimension, 689, 693–695in three dimensions, 695, 714, 737st<strong>and</strong>ing waves, 693wave number, 437, 693nwave packet, 436wave vector, k, 437wavefunction of electron in hydrogen atom, 208Weber functions Y ν (x), 607wedge product, see vector productweightof relative tensor, 964of variable, 477weight function, 555, 790Wiener–Kinchin theorem, 450WKB methods, 895–905accuracy, 902general solutions, 897phase memory, 895the Stokes phenomenon, 903work doneby <strong>for</strong>ce, 381vector representation, 220Wronskian<strong>and</strong> Green’s functions, 527<strong>for</strong> second solution of ODE, 544, 580from ODE, 532test <strong>for</strong> linear independence, 491, 532X-ray scattering, 237Yl m (θ, φ), see spherical harmonicsY ν (x), Bessel functions of second kind, 607Young’s modulus, 677, 953z, as a complex number, 84z ∗ , as complex conjugate, 89–91zero (null)matrix, 254, 255operator, 249unphysical state |∅〉, 650vector, 214, 242, 5561333

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!