11.08.2013 Views

Untitled - Cdm.unimo.it

Untitled - Cdm.unimo.it

Untitled - Cdm.unimo.it

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

PREFACE<br />

This book is devoted to the analysis of approximate solution techniques for differential<br />

equations, based on classical orthogonal polynomials. These techniques are popularly<br />

known as spectral methods. In the last few decades, there has been a growing interest<br />

in this subject. As a matter of fact, spectral methods provide a compet<strong>it</strong>ive alternative<br />

to other standard approximation techniques, for a large variety of problems. In<strong>it</strong>ial ap-<br />

plications were concerned w<strong>it</strong>h the investigation of periodic solutions of boundary value<br />

problems using trigonometric polynomials. Subsequently, the analysis was extended to<br />

algebraic polynomials. Expansions in orthogonal basis functions were preferred, due to<br />

their high accuracy and flexibil<strong>it</strong>y in computations.<br />

The aim of this book is to present a preliminary mathematical background for be-<br />

ginners who wish to study and perform numerical experiments, or who wish to improve<br />

their skill in order to tackle more specific applications. In add<strong>it</strong>ion, <strong>it</strong> furnishes a com-<br />

prehensive collection of basic formulas and theorems that are useful for implementations<br />

at any level of complex<strong>it</strong>y. We tried to maintain an elementary expos<strong>it</strong>ion so that no<br />

experience in functional analysis is required.<br />

The book is divided into thirteen chapters. In the first chapter, the reader can find<br />

defin<strong>it</strong>ions and properties relative to different families of polynomials. Orthogonal<strong>it</strong>y,<br />

zeroes, Gaussian quadrature formulas, basis transformations, and many other algebraic<br />

relations are studied in chapters 2, 3, 4. W<strong>it</strong>h the help of the short introduction to func-<br />

tional analysis given in chapter 5, a survey of results in approximation theory is provided


vi Polynomial Approximation of Differential Equations<br />

in chapter 6. Chapter 7 is devoted to the study of the discretization of derivative op-<br />

erators. The eigenvalues of the corresponding matrices are analyzed in chapter 8. The<br />

study of the differential equations begins in chapter 9. Ordinary differential equations<br />

and time-dependent partial differential equations in one space variable are examined<br />

in chapters 9 and 10 respectively. Domain-decompos<strong>it</strong>ion methods and related solu-<br />

tion techniques are considered in chapter 11. Numerical experiments on four upgraded<br />

model problems in one dimension are presented in chapter 12. Finally, suggestions for<br />

the treatment of problems in two variables are presented in the last chapter.<br />

Acknowledgments<br />

The author is grateful for the helpful scientific support of Claudio Canuto and his cr<strong>it</strong>ical<br />

revision of the entire manuscript. Useful comments and suggestions were provided by<br />

David Gottlieb for chapter 10 and Alfio Quarteroni for chapter 11.<br />

Special thanks are due to the ed<strong>it</strong>ors, Yousuff Hussaini and Gordon Erlebacher, for<br />

the uncommon effort of making the expos<strong>it</strong>ion more fluent.<br />

Final thanks go to the Ist<strong>it</strong>uto di Analisi Numerica del C.N.R., Pavia, for funding<br />

this research in part.<br />

Daniele Funaro<br />

Pavia, December 1991


CONTENTS<br />

1. Special families of polynomials<br />

1.1 Sturm-Liouville problems 1<br />

1.2 The Gamma function 2<br />

1.3 Jacobi polynomials 4<br />

1.4 Legendre polynomials 7<br />

1.5 Chebyshev polynomials 9<br />

1.6 Laguerre polynomials 12<br />

1.7 Herm<strong>it</strong>e polynomials 16<br />

2. Orthogonal<strong>it</strong>y<br />

2.1 Inner products and norms 21<br />

2.2 Orthogonal functions 23<br />

2.3 Fourier coefficients 27<br />

2.4 The projection operator 30<br />

2.5 The maximum norm 31<br />

2.6 Basis transformations 32<br />

3. Numerical integration<br />

3.1 Zeroes of orthogonal polynomials 35<br />

3.2 Lagrange polynomials 42<br />

3.3 The interpolation operators 46


viii Polynomial Approximation of Differential Equations<br />

3.4 Gauss integration formulas 47<br />

3.5 Gauss-Lobatto integration formulas 51<br />

3.6 Gauss-Radau integration formulas 54<br />

3.7 Clenshaw-Curtis integration formulas 55<br />

3.8 Discrete norms 56<br />

3.9 Discrete maximum norms 60<br />

3.10 Scaled weights 62<br />

4. Transforms<br />

4.1 Fourier transforms 65<br />

4.2 Aliasing 68<br />

4.3 Fast Fourier transform 70<br />

4.4 Other fast methods 76<br />

5. Functional spaces<br />

5.1 The Lebesgue integral 77<br />

5.2 Spaces of measurable functions 81<br />

5.3 Completeness 82<br />

5.4 Weak derivatives 84<br />

5.5 Transformation of measurable functions in R 85<br />

5.6 Sobolev spaces in R 87<br />

5.7 Sobolev spaces in intervals 90<br />

6. Results in approximation theory<br />

6.1 The problem of best approximation 93<br />

6.2 Estimates for the projection operator 96<br />

6.3 Inverse inequal<strong>it</strong>ies 104<br />

6.4 Other projection operators 107<br />

6.5 Convergence of the Gaussian formulas 111<br />

6.6 Estimates for the interpolation operator 115<br />

6.7 Laguerre and Herm<strong>it</strong>e functions 119<br />

6.8 Refinements 123


Contents ix<br />

7. Derivative Matrices<br />

7.1 Derivative matrices in the frequency space 125<br />

7.2 Derivative matrices in the physical space 129<br />

7.3 Boundary cond<strong>it</strong>ions in the frequency space 135<br />

7.4 Boundary cond<strong>it</strong>ions in the physical space 138<br />

7.5 Derivatives of scaled functions 146<br />

7.6 Numerical solution 148<br />

8. Eigenvalue analysis<br />

8.1 Eigenvalues of first derivative operators 151<br />

8.2 Eigenvalues of higher-order operators 157<br />

8.3 Cond<strong>it</strong>ion number 165<br />

8.4 Precond<strong>it</strong>ioners for second-order operators 168<br />

8.5 Precond<strong>it</strong>ioners for first-order operators 172<br />

8.6 Convergence of eigenvalues 174<br />

8.7 Multigrid method 179<br />

9. Ordinary differential equations<br />

9.1 General considerations 181<br />

9.2 Approximation of linear equations 183<br />

9.3 The weak formulation 189<br />

9.4 Approximation of problems in the weak form 197<br />

9.5 Approximation in unbounded domains 207<br />

9.6 Other techniques 210<br />

9.7 Boundary layers 211<br />

9.8 Nonlinear equations 214<br />

9.9 Systems of equations 217<br />

9.10 Integral equations 218<br />

10. Time-dependent problems<br />

10.1 The Gronwall inequal<strong>it</strong>y 221<br />

10.2 Approximation of the heat equation 222


x Polynomial Approximation of Differential Equations<br />

10.3 Approximation of linear first-order problems 229<br />

10.4 Nonlinear time-dependent problems 235<br />

10.5 Approximation of the wave equation 237<br />

10.6 Time discretization 243<br />

11. Domain-decompos<strong>it</strong>ion methods<br />

11.1 Introductory remarks 249<br />

11.2 Non-overlapping multidomain methods 250<br />

11.3 Solution techniques 260<br />

11.4 Overlapping multidomain methods 263<br />

12. Examples<br />

12.1 A free boundary problem 265<br />

12.2 An example in an unbounded domain 271<br />

12.3 The nonlinear Schrödinger equation 274<br />

12.4 Zeroes of Bessel functions 278<br />

13. An example in two dimensions<br />

13.1 Poisson’s equation 281<br />

13.2 Approximation by the collocation method 282<br />

13.3 Hints for the implementation 287<br />

13.4 The incompressible Navier-Stokes equations 290<br />

References 293<br />

Index 305


1<br />

SPECIAL FAMILIES<br />

OF POLYNOMIALS<br />

Approximating functions in spectral methods are related to polynomial solutions of<br />

eigenvalue problems in ordinary differential equations, known as Sturm-Liouville prob-<br />

lems. These originate from applying the method of separation of variables in the analysis<br />

of boundary-value problems. We outline both basic and remarkable properties of the<br />

most commonly used families of polynomials of this kind.<br />

1.1 Sturm-Liouville problems<br />

Let I denote an open interval in R. We consider the continuous functions a : Ī → R,<br />

b : I → R, w : I → R, satisfying a ≥ 0 in Ī and w > 0 in I. We are concerned w<strong>it</strong>h<br />

the solutions (λ,u) to the following eigenvalue problem:<br />

(1.1.1) −(au ′ ) ′ + bu = λw u in I , λ ∈ R.<br />

A large variety of boundary cond<strong>it</strong>ions can be assumed in order to determine uniquely,<br />

up to a constant factor, the eigenfunction u. Problem (1.1.1) is said to be regular if<br />

a > 0 in Ī. When a vanishes at least at one point in Ī, then we have a singular problem.<br />

The l<strong>it</strong>erature about this subject gives plenty of results. For a general theory, we can<br />

refer for instance to courant and hilbert(1953), yosida(1960), brauer and nohel


2 Polynomial Approximation of Differential Equations<br />

(1967), dettman(1969). We only remark that, in most of the cases, one can show that<br />

the set of eigenvalues form a divergent sequence of real pos<strong>it</strong>ive numbers. Many families<br />

of eigenfunctions have been widely studied. Among these, Bessel functions are the most<br />

representative (see section 12.4). However, we are mainly concerned w<strong>it</strong>h polynomial<br />

solutions of (1.1.1). In particular, we are interested in introducing and characterizing<br />

sequences {(λn, un)} n∈N of solutions to (1.1.1) w<strong>it</strong>h λn > 0, lim<br />

n→∞ λn = + ∞, where<br />

the un’s are polynomials of degree n, su<strong>it</strong>ably normalized. Many examples will be given<br />

in the following. Here we state a first general result, which will be proven in section 2.2.<br />

Theorem 1.1.1 - If {un}n∈N is a sequence of solutions to (1.1.1), where un is a poly-<br />

nomial of degree n, then <strong>it</strong> is possible to find three real sequences {ρn}, {σn}, {τn}, n ≥ 2<br />

such that<br />

(1.1.2) un (x) = (ρn x + σn) un−1 (x) + τn un−2 (x), ∀n ≥ 2, ∀x ∈ I.<br />

According to (1.1.2), we define ρ1, σ1 ∈ R, such that u1(x) = (ρ1x + σ1)u0(x).<br />

Theorem 1.1.1 allows us to recursively compute the n th polynomial at a given point<br />

x ∈ I, starting from the values u0(x) and u1(x). By differentiating the expression<br />

(1.1.2), we get a similar relation for the derivative, i.e.<br />

(1.1.3) u ′ n (x) = (ρn x + σn) u ′ n−1 (x) + ρn un−1 (x) + τn u ′ n−2 (x), ∀n ≥ 2, ∀x ∈ I.<br />

The simultaneous evaluation of (1.1.2) and (1.1.3) gives u ′ n at the point x. Evaluation<br />

of higher order derivatives can be clearly carried out in the same way.<br />

1.2 The Gamma function<br />

Before continuing our analysis of polynomial eigenfunctions, <strong>it</strong> is convenient to review<br />

some properties of the well-known Euler Gamma function which, for any real pos<strong>it</strong>ive


Special Families of Polynomials 3<br />

number x, is defined as<br />

Γ(x) :=<br />

<br />

+∞<br />

0<br />

t x−1 e −t dt.<br />

Proving the following functional equation is straightforward after integration by parts:<br />

(1.2.1) Γ(x + 1) = xΓ(x) , x > 0.<br />

In particular, when n is an integer, we have a fundamental relation obtained from (1.2.1)<br />

by induction, i.e. Γ(n + 1) = n!.<br />

Another important relation is<br />

(1.2.2) Γ(x) Γ(1 − x) = π<br />

, 0 < x < 1.<br />

sinπ x<br />

From (1.2.2) one can easily obtain Γ √ 1<br />

2 = π. A useful equal<strong>it</strong>y, which can be deduced<br />

from the defin<strong>it</strong>ion of the Beta function (see for instance luke (1969), Vol.1), is<br />

(1.2.3)<br />

1<br />

−1<br />

(1 + t) x−1 (1 − t) y−1 x+y−1 Γ(x) Γ(y)<br />

dt = 2 , x > 0, y > 0.<br />

Γ(x + y)<br />

The binomial coefficients are generalized according to<br />

(1.2.4)<br />

<br />

x<br />

:=<br />

k<br />

Γ(x + 1)<br />

, k ∈ N, x > k − 1.<br />

k! Γ(x − k + 1)<br />

The formulas considered here, as well as many other properties of the Gamma<br />

function are widely discussed in luke(1969), Vol.1; srivastava and manocha(1984),<br />

hochstadt(1971).<br />

The numerical evaluation of the Gamma function is generally performed by reduc-<br />

ing the computation shifting backwards to the interval ]0,1] using (1.2.1). The interval<br />

can be further reduced to ] 1<br />

2 , 1] by taking into account (1.2.2). Finally, su<strong>it</strong>able series<br />

expansions centered about 1 can be implemented. For other theoretical and compu-<br />

tational aspects we can refer for instance to luke(1969), Vol.2. We only mention the<br />

following expansion:


4 Polynomial Approximation of Differential Equations<br />

(1.2.5) Γ(x) =<br />

where<br />

τ0 :=<br />

Hk(x) :=<br />

<br />

x + 9<br />

1 x− 2<br />

2<br />

2<br />

11 , τk := 2(−1) k k<br />

e 1−x<br />

<br />

τ0 +<br />

∞<br />

k=1<br />

(x − 1) (x − 2) · · · (x − k)<br />

x(x + 1) · · · (x + k − 1)<br />

k<br />

m=0<br />

(k + m − 1)!<br />

m!(k − m)!<br />

τk Hk (x)<br />

<br />

, k ≥ 1 ,<br />

,<br />

(−1) m em (m + 11 , k ≥ 1.<br />

2<br />

)m+1/2<br />

This is basically the expression given in luke(1969), Vol.1, page 30. More theoretical<br />

details and tabulated values are contained in that book.<br />

1.3 Jacobi polynomials<br />

We introduce the first family of polynomial solutions to (1.1.1) called Jacobi polynomi-<br />

als. They depend on two parameters α, β ∈ R w<strong>it</strong>h α > −1, β > −1. As we will see<br />

in the following, an appropriate choice of these parameters leads to other well-known<br />

families, such as Legendre or Chebyshev polynomials.<br />

Let us set I =] − 1,1[ and let us take a, b, w in (1.1.1) such that<br />

a(x) = (1 − x) α+1 (1 + x) β+1 , ∀x ∈ Ī,<br />

b(x) = 0 , w(x) = (1 − x) α (1 + x) β , ∀x ∈ I,<br />

α > −1, β > −1.<br />

After simplification, this choice yields the singular eigenvalue equation<br />

(1.3.1) − 1 − x 2 u ′′ + (α + β + 2)x + α − β u ′ = λu.


Special Families of Polynomials 5<br />

By applying the Frobenius method (see for instance dettman(1969)), we get the fol-<br />

lowing result.<br />

Theorem 1.3.1 - The solution to (1.3.1) is a polynomial of degree n only when<br />

λ = n(n + α + β + 1), n ∈ N.<br />

We shall define the n-degree Jacobi polynomial P (α,β)<br />

n<br />

(1.3.1) normalized by<br />

(1.3.2) P (α,β)<br />

n (1) :=<br />

or equivalently by<br />

<br />

n + α<br />

=<br />

n<br />

Γ(n + α + 1)<br />

(1.3.3) P (α,β)<br />

n (−1) := (−1) n<br />

n! Γ(α + 1)<br />

to be the unique solution of<br />

, n ∈ N,<br />

<br />

n + β<br />

, n ∈ N.<br />

n<br />

We observe that no boundary cond<strong>it</strong>ions are imposed in (1.3.1). These are replaced<br />

by requiring the solutions to be polynomials. Cond<strong>it</strong>ion (1.3.2) (or (1.3.3)) is only<br />

imposed to select a unique eigenfunction, otherwise defined to w<strong>it</strong>hin a multiplicative<br />

constant.<br />

Many theorems and properties for this family of polynomials are well-known. Here<br />

we just collect some basic results and we refer to szegö(1939), luke(1969), askey(1975),<br />

srivastava and manocha(1984) for a detailed essay. First we give the Rodrigues’ for-<br />

mula:<br />

(1.3.4) (1 − x) α (1 + x) β P (α,β)<br />

n (x) = (−1)n<br />

2n n!<br />

A more explic<strong>it</strong> form is<br />

(1.3.5) P (α,β)<br />

n (x) = 2 −n<br />

=<br />

Γ(2n + α + β + 1)<br />

2 n n! Γ(n + α + β + 1)<br />

n<br />

<br />

n + α n + β<br />

k n − k<br />

k=0<br />

<br />

x n +<br />

dn dxn n+α n+β<br />

(1 − x) (1 + x) , n ∈ N.<br />

(x − 1) n−k (x + 1) k<br />

(α − β) n<br />

2n + α + β xn−1 <br />

+ · · · , n ∈ N.


6 Polynomial Approximation of Differential Equations<br />

Both (1.3.4) and (1.3.5) can be checked by direct subst<strong>it</strong>ution in (1.3.1). A lot of<br />

formulas relate Jacobi polynomials corresponding to different choices of the pair of<br />

parameters (α, β). We show one of the most representative, i.e.<br />

(1.3.6)<br />

d<br />

<br />

dx<br />

P (α,β)<br />

n<br />

<br />

= 1<br />

2<br />

By (1.3.4) or (1.3.5) <strong>it</strong> is clear that<br />

(α, β)<br />

(1.3.7) P<br />

0 (x) = 1 , P<br />

(n + α + β + 1) P (α+1, β+1)<br />

n−1 , n ≥ 1.<br />

(α, β)<br />

1<br />

(x) = 1<br />

1<br />

(α + β + 2)x + (α − β).<br />

2 2<br />

Since (1.3.4) and (1.3.5) are qu<strong>it</strong>e unpracticable, higher degree polynomials can be<br />

determined using theorem 1.1.1. More precisely, we have<br />

(1.3.8) ρn =<br />

τn = −<br />

σn =<br />

(2n + α + β) (2n + α + β − 1)<br />

,<br />

2n(n + α + β)<br />

(α 2 − β 2 ) (2n + α + β − 1)<br />

2n (n + α + β) (2n + α + β − 2) ,<br />

Moreover, one has: ρ1 = 1<br />

2 (α + β + 2), σ1 = 1<br />

2<br />

using (1.1.3) or (1.3.6).<br />

(n + α − 1) (n + β − 1) (2n + α + β)<br />

, n ≥ 2.<br />

n (n + α + β) (2n + α + β − 2)<br />

(α − β). Derivatives can be recovered<br />

Finally, we present the following estimate in the interval Ī (see szegö (1939)):<br />

(1.3.9) max<br />

−1≤x≤1<br />

|P (α,β)<br />

n<br />

<br />

(x) | = max |P (α,β)<br />

<br />

n (±1)| = max<br />

n + α<br />

n<br />

<br />

,<br />

n + β<br />

n<br />

<br />

,<br />

n ∈ N.<br />

Further properties can be proven for the so called ultraspherical (or Gegenbauer)<br />

polynomials. These are Jacobi polynomials where α = β. We soon see two celebrated<br />

examples.


Special Families of Polynomials 7<br />

1.4 Legendre polynomials<br />

Legendre polynomials are Jacobi ultraspherical polynomials w<strong>it</strong>h α = β = 0. To simplify<br />

the notation <strong>it</strong> is standard to set Pn := P (0,0)<br />

n . We now review the basic properties.<br />

According to theorem 1.3.1, we have the differential equation<br />

(1.4.1) (1 − x 2 ) P ′′<br />

n − 2xP ′ n + n(n + 1)Pn = 0 , n ∈ N.<br />

Cond<strong>it</strong>ions (1.3.2), (1.3.3) give respectively Pn (1) = 1, Pn (−1) = (−1) n , n ∈ N. The<br />

recursion formula is<br />

(1.4.2) Pn(x) =<br />

2n − 1<br />

n<br />

xPn−1(x) −<br />

n − 1<br />

n Pn−2(x), x ∈ Ī, n ≥ 2.<br />

Figure 1.4.1 - Legendre polynomials Figure 1.4.2 - The eleventh Legendre<br />

for 1 ≤ n ≤ 6. polynomial.<br />

It is easy to check that Pn is an even (odd) function if and only if n is even (odd).<br />

Moreover, from (1.4.1) one gets<br />

(1.4.3) P ′ n(±1) = ±<br />

n(n + 1)<br />

2<br />

(±1) n , n ∈ N.


8 Polynomial Approximation of Differential Equations<br />

Furthermore, (1.3.9) leads to<br />

(1.4.4) |Pn(x)| ≤ 1 , |x| ≤ 1, n ∈ N.<br />

Combining (1.3.6) and (1.3.9), we can also estimate the derivative according to<br />

(1.4.5) |P ′ n(x)| ≤ 1<br />

2n(n + 1) , |x| ≤ 1 , n ∈ N.<br />

Another useful relation can be easily established by taking x = 0 in (1.4.2), i.e.<br />

⎧<br />

⎨0<br />

if n is odd,<br />

(1.4.6) Pn(0) =<br />

⎩<br />

n! 2−n <br />

n −2<br />

2 ! if n is even.<br />

Therefore, one deduces that lim<br />

n→+ ∞ Pn(0) = 0.<br />

In Figure 1.4.1 the polynomials Pn, 1 ≤ n ≤ 6, are plotted, while Figure 1.4.2<br />

shows the behavior of P11.<br />

In order to give some general ideas of the behavior of Legendre polynomials, we<br />

state two other results (see szegö(1939) and jackson(1930) respectively).<br />

Theorem 1.4.1. - For any n ≥ 5, when x increases from 0 to 1, the successive relative<br />

maximum values of |Pn(x)| also increase.<br />

Theorem 1.4.2. - For any n ∈ N and any x ∈ Ī, one has<br />

(1.4.7) Pn(x) = 1<br />

π<br />

π<br />

0<br />

<br />

x + i 1 − x2 n cos θ<br />

Though the right hand side in (1.4.7) contains the imaginary un<strong>it</strong>y (i.e., i = √ −1), the<br />

global integral is real. Taking the absolute value on both sides of (1.4.7), we can prove<br />

the estimate:<br />

dθ.


Special Families of Polynomials 9<br />

(1.4.8) |Pn(x)| ≤ 1<br />

<br />

π<br />

≤ 1<br />

<br />

π<br />

0<br />

π<br />

0<br />

π<br />

x 2 + (1 − x 2 ) cos 2 θ n/2 dθ<br />

2 2 2 1 + x<br />

x + (1 − x ) cos θ dθ = 2<br />

, x ∈ Ī, n ≥ 2.<br />

2<br />

This shows that the Legendre polynomials in<br />

parabolas y = − 1<br />

2 (1 + x2 ) and y = 1<br />

2 (1 + x2 ).<br />

1.5 Chebyshev polynomials<br />

Ī are uniformly bounded between the<br />

The Chebyshev polynomials (of the first kind) are related to the ultraspherical polyno-<br />

mials w<strong>it</strong>h α = β = −1 2 . In fact, they are defined by<br />

(− 1<br />

,− 1<br />

2 )<br />

2 (1.5.1) Tn := δn P n , n ∈ N,<br />

where<br />

δn := (n!2n ) 2<br />

(2n)! = n! √ π<br />

Γ(n + 1 =<br />

)<br />

2<br />

1−1<br />

n − 2 .<br />

n<br />

Consequently, they are solutions of the Sturm-Liouville problem<br />

(1.5.2) (1 − x 2 ) T ′′<br />

n − xT ′ n + n 2 Tn = 0 , n ∈ N.<br />

By taking α = β = − 1<br />

2<br />

formula<br />

in (1.3.8), after appropriate scaling, we arrive at the recursion<br />

(1.5.3) Tn(x) = 2xTn−1(x) − Tn−2(x) , x ∈<br />

where T0(x) = 1 and T1(x) = x.<br />

Ī , n ≥ 2,


10 Polynomial Approximation of Differential Equations<br />

In general, we obtain Tn(±1) = (±1) n , n ∈ N. Evaluating at the points x = ±1,<br />

(1.5.2) yields<br />

(1.5.4) T ′ n(±1) = ±(±1) n n 2 , n ∈ N.<br />

Moreover, Tn is even (odd) if and only if n is even (odd). An explic<strong>it</strong> expression for the<br />

Chebyshev polynomial of degree n is<br />

(1.5.5) Tn(x) =<br />

[n/2] <br />

k=0<br />

⎛<br />

⎝(−1) k<br />

[n/2] <br />

m=k<br />

<br />

n m<br />

2m k<br />

⎞<br />

⎠ x n−2k<br />

= 2 n−1 x n − n2 n−3 x n−2 + 1<br />

2 n(n − 3) 2n−5 x n−4 + · · · , x ∈<br />

In (1.5.5), [•] denotes the integer part of •.<br />

The most remarkable characterization is given by the simple relation<br />

(1.5.6) Tn (cos θ) = cos nθ, θ ∈ [0, π], n ∈ N.<br />

Ī, n ∈ N.<br />

The above expression relates algebraic and trigonometric polynomials. Such an impor-<br />

tant peculiar<strong>it</strong>y can be proven by noting that<br />

(1.5.7)<br />

(1.5.8)<br />

dTn<br />

dx<br />

(cos θ) = n sinnθ<br />

sin θ ,<br />

d2 Tn n sinnθ cos θ<br />

(cos θ) =<br />

dx2 sin 3 θ<br />

− n2 cos nθ<br />

sin 3 ,<br />

θ<br />

where x = cos θ , θ ∈ [0,π], and n ∈ N. Thus, (1.5.2) is satisfied by replacing x by<br />

cos θ.<br />

A lot of properties are direct consequence of (1.5.6). In particular we have |Tn(x)| ≤ 1,<br />

∀n ∈ N, ∀x ∈ Ī. Moreover, Tn vanishes n times in Ī, therefore the derivative T ′ n<br />

vanishes n − 1 times in Ī. The function |Tn| attains the maximum value of 1, n + 1<br />

times in Ī. We give the plot in Figure 1.5.1 of the Tn’s for 1 ≤ n ≤ 6. In add<strong>it</strong>ion, T11<br />

is plotted in Figure 1.5.2.


Special Families of Polynomials 11<br />

Figure 1.5.1 - Chebyshev polynomials Figure 1.5.2 - The eleventh Chebyshev<br />

for 1 ≤ n ≤ 6. polynomial.<br />

Other simple relations can be easily established:<br />

(1.5.9) Tn(x) = 1 i n θ −i n θ<br />

e + e<br />

2<br />

= 1<br />

<br />

x +<br />

2<br />

x2 n <br />

− 1 + x − x2 n − 1 ,<br />

(1.5.10) Tn = 1<br />

′ T n+1<br />

2 n + 1 − T ′ <br />

n−1<br />

n − 1<br />

<br />

(1.5.11) Tm Tn(x) = Tmn (x) , ∀x ∈<br />

, n ≥ 2,<br />

We refer to luke(1969) and rivlin(1974) for more details.<br />

Ī, ∀n, m ∈ N.<br />

x = cos θ ∈ Ī, n ∈ N,<br />

Another family of polynomials, known as Chebyshev polynomials of the second kind<br />

is defined as<br />

(1.5.12) Un := 1<br />

n + 1 T ′ n+1 , n ∈ N.


12 Polynomial Approximation of Differential Equations<br />

These are also ultraspherical polynomials. By virtue of (1.3.6) and (1.5.1), we have<br />

2 (1.5.13) Un = δn+1 P n , n ∈ N.<br />

( 1<br />

, 1<br />

2 )<br />

The Un’s satisfy similar properties to those of Chebyshev polynomials of the first kind.<br />

1.6 Laguerre polynomials<br />

In this section, we introduce another family of polynomial solutions of the Sturm-<br />

Liouville problem. Let α > −1 and I =]0,+∞[, then define the coefficients in (1.1.1)<br />

to be<br />

a(x) = x α+1 e −x , ∀x ∈ Ī,<br />

b(x) = 0, w(x) = x α e −x , ∀x ∈ I.<br />

This choice leads to the eigenvalue problem<br />

(1.6.1) xu ′′ + (α + 1 − x) u ′ + λu = 0.<br />

This adm<strong>it</strong>s polynomial solutions only if λ = n, n ∈ N. Therefore, we define the<br />

n-degree Laguerre polynonial L (α)<br />

n<br />

(1.6.1), satisfying the cond<strong>it</strong>ion<br />

(1.6.2) L (α)<br />

n (0) :=<br />

We have the Rodrigues’ formula<br />

(1.6.3) e −x x α L (α)<br />

n (x) = 1<br />

n!<br />

The following expression also holds:<br />

(1.6.4) L (α)<br />

n (x) =<br />

n<br />

k=0<br />

to be the unique solution of the singular problem<br />

n + α<br />

n<br />

<br />

, n ∈ N, α > −1.<br />

dn dxn −x n+α<br />

e x , n ∈ N, x ∈ I.<br />

k n + α (−1)<br />

n − k k !<br />

x k , n ∈ N, x ∈ Ī.


Special Families of Polynomials 13<br />

By virtue of theorem 1.1.1, we have the three-term recursion formula<br />

(1.6.5) L (α)<br />

n (x) =<br />

where L (α)<br />

0<br />

2n + α − 1 − x<br />

n<br />

(x) = 1 and L(α) 1 (x) = 1 + α − x.<br />

L (α) n + α − 1<br />

n−1 (x) −<br />

n<br />

L (α)<br />

n−2 (x), ∀n ≥ 2,<br />

Various equations relate Laguerre polynomials corresponding to different values of<br />

the parameter α. Some examples are (see also szegö (1939))<br />

(1.6.6)<br />

d<br />

dx L(α) n+1 = −L(α+1) n , n ∈ N, α > −1,<br />

(1.6.7) L (α)<br />

n+1 = L(α+1) n+1 − L(α+1) n , n ∈ N, α > −1.<br />

From (1.6.7) one obtains<br />

(1.6.8) L (α+1)<br />

n<br />

=<br />

n<br />

k=0<br />

L (α)<br />

k , n ∈ N, α > −1.<br />

An asymptotic formula relates Laguerre and Jacobi polynomials, namely:<br />

(1.6.9) L (α)<br />

n (x) = lim<br />

β→+∞<br />

<br />

P (α,β)<br />

n<br />

<br />

1 − 2x<br />

β<br />

<br />

, ∀x ∈ [0,+∞[.<br />

This can be checked w<strong>it</strong>h the help of the differential equations (1.3.1) and (1.6.1).<br />

A statement similar to that of theorem 1.4.1 also holds (see szegö(1939), p.171).<br />

Theorem 1.6.1 - For any α > −1 and n ≥ 4, the successive values of the relative<br />

maxima of |L (α)<br />

n (x)| are decreasing when x < α+ 1<br />

2<br />

and increasing when x > α+ 1<br />

2 .<br />

The plots of figures 1.6.1 and 1.6.2 show respectively L (0)<br />

n , 1 ≤ n ≤ 6, and L (0)<br />

7 .<br />

The size of the window is [0,20] × [−600,600] in figure 1.6.1 and [0,20] × [−900,900] in<br />

figure 1.6.2. When n grows, graphical representations of Laguerre polynomials are dif-<br />

ficult to obtain. An in<strong>it</strong>ial gentle behavior explodes to severe oscillations, for increasing<br />

values of x.


14 Polynomial Approximation of Differential Equations<br />

Figure 1.6.1 - Laguerre polynomials Figure 1.6.2 - The Laguerre<br />

for α = 0 and 1 ≤ n ≤ 6. polynomial L (0)<br />

7 .<br />

Further insight into the asymptotic behavior can be given. We recall two results<br />

(see szegö(1939), p.171 and p.234). The first one shows the behavior when x → +∞.<br />

In the second one, the case n → +∞ is considered. It is convenient to define Q (α)<br />

n (x) :=<br />

e −x/2 x (α+1)/2 L (α)<br />

n (x).<br />

Theorem 1.6.2 - For any α > −1 and n ≥ 2, the values of the relative maxima of<br />

|Q (α)<br />

n (x)| form an increasing sequence when x ∈]x0,+∞[, where<br />

x0 := 0 if α ≤ 1, x0 :=<br />

α 2 − 1<br />

2n + α + 1<br />

if α > 1.<br />

Theorem 1.6.3 - Let α > −1, then for any γ > 0 and 0 < η < 4 we can find two<br />

pos<strong>it</strong>ive constants C1, C2 such that<br />

(1.6.10) max<br />

x∈[γ,(4−η)n]<br />

(1.6.11) max<br />

x∈[γ,+∞[<br />

|Q (α)<br />

n (x)| ≈ C1 n α/2 ,<br />

|Q (α)<br />

n (x)| ≈ C2 6√ n n α/2 .


Special Families of Polynomials 15<br />

In order to allow more flexible numerical computations, <strong>it</strong> is useful to introduce<br />

some su<strong>it</strong>able scaling functions S (α)<br />

n : Ī → R and to define<br />

(1.6.12)<br />

ˆ L (α)<br />

n := S (α)<br />

n L (α)<br />

n , n ∈ N, α > −1.<br />

The aim is to avoid ill-cond<strong>it</strong>ioned operations when evaluating point values of Laguerre<br />

polynomials. According to funaro(1990a) an effective choice of S (α)<br />

n is<br />

(1.6.13) S (α)<br />

0 (x) := 1, S (α)<br />

n (x) :=<br />

n n<br />

+ α <br />

n<br />

k=1<br />

<br />

1 + x<br />

<br />

4k<br />

−1 , n ≥ 1.<br />

We will refer to ˆ L (α)<br />

n , n ∈ N, as the family of scaled Laguerre functions. It is clear that<br />

these are not polynomials. Plots of ˆ L (0)<br />

n , 1 ≤ n ≤ 12, are given in figure 1.6.3. Now<br />

the window size is [0,50] × [−300,300].<br />

Figure 1.6.3 - Scaled Laguerre functions for α = 0 and 1 ≤ n ≤ 12.


16 Polynomial Approximation of Differential Equations<br />

It is evident that the scaled Laguerre functions have a milder behavior. We note that<br />

ˆL (α)<br />

n (0) = 1, ∀n ∈ N. More details are given in funaro(1990a). By subst<strong>it</strong>ution in<br />

(1.6.5), after simplification, one obtains ∀n ≥ 2 the recursion formula<br />

(1.6.14)<br />

ˆ L (α)<br />

n (x) =<br />

w<strong>it</strong>h ˆ L (α)<br />

0 (x) = 1 and ˆ L (α)<br />

1 (x) = 4(α+1−x)<br />

(α+1)(x+4) .<br />

Moreover, for the derivatives one has ∀n ≥ 2<br />

d<br />

(1.6.15)<br />

dx ˆ L (α)<br />

<br />

4n<br />

n (x) =<br />

(n + α)(4n + x)<br />

6n + α − 1<br />

−<br />

4n + x ˆ L (α)<br />

<br />

4(n − 1)2<br />

n−1 (x) +<br />

4n + x − 4<br />

w<strong>it</strong>h d<br />

dx ˆ L (α)<br />

0 (x) = 0 and d<br />

<br />

4n<br />

(2n + α − 1 − x)<br />

(n + α)(4n + x)<br />

ˆ L (α)<br />

n−1 (x)<br />

dx ˆ L (α)<br />

1<br />

−<br />

4(n − 1)2<br />

4n + x − 4 ˆ L (α)<br />

n−2 (x)<br />

<br />

,<br />

(2n + α − 1 − x) d<br />

dx ˆ L (α)<br />

n−1 (x)<br />

2(4n + x − 2)<br />

(4n + x)(4n + x − 4) ˆ L (α) d<br />

n−2 (x) −<br />

dx ˆ L (α)<br />

n−2 (x)<br />

(x) = − α+5<br />

α+1<br />

4<br />

(x+4) 2 .<br />

Applications will be examined in the following chapters (see sections 3.10 and 7.5).<br />

1.7 Herm<strong>it</strong>e polynomials<br />

The set of Herm<strong>it</strong>e polynomials Hn, n ∈ N, is the last family we consider. Following<br />

the notations adopted in szegö(1939) or courant and hilbert(1953), the Hn’s are<br />

solutions of the non singular Sturm-Liouville problem obtained by setting in (1.1.1):<br />

<br />

,


Special Families of Polynomials 17<br />

a(x) = w(x) = e −x2<br />

, b(x) = 0, ∀x ∈ I ≡ R.<br />

Therefore, one easily gets the differential equation<br />

(1.7.1) H ′′<br />

n(x) − 2xH ′ n(x) + 2nHn(x) = 0, n ∈ N, x ∈ R.<br />

The normalizing cond<strong>it</strong>ion is<br />

(1.7.2) Hn(0) := (−1) n/2 n!<br />

(n/2)!<br />

(1.7.3) H ′ n(0) := (−1) (n−1)/2 (n + 1)!<br />

((n + 1)/2)!<br />

if n is even,<br />

if n is odd.<br />

The n th degree polynomial Hn is an even or odd function according to the par<strong>it</strong>y of n.<br />

As usual, we have the Rodrigues’ formula<br />

(1.7.4) Hn(x) = (−1) n dn x2<br />

e e−x2,<br />

n ∈ N, x ∈ R.<br />

dxn More explic<strong>it</strong>ly, we can wr<strong>it</strong>e<br />

(1.7.5) Hn(x) = n!<br />

[n/2] <br />

m=0<br />

(−1) m<br />

m!<br />

(2x) n−2m<br />

(n − 2m)!<br />

= 2 n x n − n(n − 1)2 n−2 x n−2 + n(n − 1)(n − 2)(n − 3)2 n−5 x n−4 − · · ·,<br />

The corresponding three-term recursion formula is<br />

(1.7.6) Hn(x) = 2xHn−1(x) − 2(n − 1)Hn−2(x), ∀n ≥ 2,<br />

where H0(x) = 1 and H1(x) = 2x.<br />

By differentiating (1.7.4) we also get<br />

(1.7.7) H ′ n(x) = 2xHn(x) − Hn+1(x), n ∈ N, x ∈ R.<br />

n ∈ N, x ∈ R.


18 Polynomial Approximation of Differential Equations<br />

Together w<strong>it</strong>h (1.7.6), we obtain a very simple relation to evaluate the derivatives, i.e.<br />

(1.7.8) H ′ n(x) = 2nHn−1(x), n ≥ 1, x ∈ R.<br />

We give the plots of Herm<strong>it</strong>e polynomials in figures 1.7.1 and 1.7.2. The sizes of<br />

the windows are respectively [−5,5] × [−900,900] and [−5,5] × [−450000,450000].<br />

Figure 1.7.1 - Herm<strong>it</strong>e polynomials Figure 1.7.2 - The ninth Herm<strong>it</strong>e<br />

for 1 ≤ n ≤ 6. polynomial.<br />

It is worthwhile to mention the following result.<br />

Theorem 1.7.1 - For any n ≥ 4, the successive values of the relative maxima of<br />

|Hn(x)| are increasing for x ≥ 0.<br />

Other asymptotic properties can be found in szegö(1939).<br />

The Herm<strong>it</strong>e polynomials can be expressed in term of the Laguerre polynomials accord-<br />

ing to


Special Families of Polynomials 19<br />

(1.7.9) Hn(x) = (−1) n/2 2 n (n/2)! L (−1/2)<br />

n/2 (x 2 ), if n is even,<br />

(1.7.10) Hn(x) = (−1) (n−1)/2 2 n ((n − 1)/2)! x L (1/2)<br />

(n−1)/2 (x2 ), if n is odd.<br />

The formulas above are easily checked w<strong>it</strong>h the help of (1.6.1)-(1.6.2) and (1.7.1)-(1.7.2)-<br />

(1.7.3).<br />

Inspired by (1.7.9) and (1.7.10), recalling the defin<strong>it</strong>ion (1.6.12)-(1.6.13), <strong>it</strong> is nat-<br />

ural to define the scaled Herm<strong>it</strong>e functions by<br />

(1.7.11)<br />

(1.7.12)<br />

Consequently<br />

ˆ Hn(x) := ˆ L (−1/2)<br />

n/2 (x 2 ), if n is even,<br />

ˆ Hn(x) := x ˆ L (1/2)<br />

(n−1)/2 (x2 ), if n is odd.<br />

(1.7.13) ˆ Hn(0) = 1, if n is even,<br />

(1.7.14) ˆ H ′ n(0) = 1, if n is odd.<br />

Derivatives can be computed via (1.6.14) and (1.6.15).<br />

Finally we show the plots of ˆ Hn, 1 ≤ n ≤ 9, in figure 1.7.3. The window measures<br />

[−5,5] × [−3,3].


20 Polynomial Approximation of Differential Equations<br />

Figure 1.7.3 - Scaled Herm<strong>it</strong>e functions for 1 ≤ n ≤ 9.<br />

* * * * * * * * * * * *<br />

Our survey is now concluded. Of course, other families of polynomials could be consid-<br />

ered. However, the ones presented here are those actually used in spectral methods.


2<br />

ORTHOGONALITY<br />

Inner products and orthogonal functions are fundamental concepts in approximation<br />

theory. All the families introduced in the first chapter form an orthogonal set of func-<br />

tions w<strong>it</strong>h respect to a su<strong>it</strong>able inner product. This property will reveal many other<br />

interesting features.<br />

2.1 Inner products and norms<br />

In the following, X will denote a real vector space, i.e. a set in which add<strong>it</strong>ion and scalar<br />

multiplication are defined w<strong>it</strong>h the usual basic properties (for the unfamiliar reader a<br />

wide bibliography is available; we suggest, for instance, hoffman and kunze(1971),<br />

lowenthal(1975)).<br />

An inner product (· , ·) : X × X → R is a bilinear application. This means that,<br />

for any couple of vectors in X, their inner product is a real number and the application<br />

is linear for each one of the two variables. Moreover, we require symmetry, i.e.<br />

(2.1.1) (u,v) = (v,u), ∀u,v ∈ X,<br />

and the application has to be pos<strong>it</strong>ive-defin<strong>it</strong>e, i.e. (here 0 is the zero of X):<br />

(2.1.2) (u,u) ≥ 0, ∀u ∈ X, (u,u) = 0 ⇐⇒ u ≡ 0.


22 Polynomial Approximation of Differential Equations<br />

A norm · : X → R + is a real pos<strong>it</strong>ive function in X w<strong>it</strong>h the properties:<br />

(2.1.3) u ≥ 0, ∀u ∈ X, u = 0 ⇐⇒ u ≡ 0,<br />

(2.1.4) λu = |λ| u, ∀u ∈ X, ∀λ ∈ R,<br />

(2.1.5) u + v ≤ u + v, ∀u,v ∈ X (triangle inequal<strong>it</strong>y).<br />

Inner products and norms are, in general, independent concepts. Nevertheless, whenever<br />

an inner product is available in X, then a norm is automatically defined by setting<br />

(2.1.6) u := (u,u), ∀u ∈ X.<br />

Checking that (2.1.6) gives actually a norm is an easy exercise. In particular (2.1.5) is<br />

a byproduct of the well-known Schwarz inequal<strong>it</strong>y<br />

(2.1.7) |(u,v)| ≤ u v, ∀u,v ∈ X.<br />

Detailed proofs of (2.1.7) and of many other properties of inner products and norms are<br />

widely available in all the basic texts of linear algebra.<br />

We give an example. Let X = C0 ( Ī) be the linear space of continuous functions<br />

in the interval Ī. Let w : I → R be a continuous integrable function satisfying w > 0.<br />

Then, when I is bounded, an inner product (· , ·)w and <strong>it</strong>s corresponding norm · w<br />

are defined by<br />

(2.1.8) (u,v)w :=<br />

(2.1.9) uw :=<br />

<br />

<br />

I<br />

I<br />

uv w dx, ∀u,v ∈ C 0 ( Ī),<br />

u 2 1<br />

w dx<br />

2<br />

, ∀u ∈ C 0 ( Ī).


Orthogonal<strong>it</strong>y 23<br />

The function w is called weight function. If I is not bounded, we just have to be careful<br />

that the integrals in (2.1.8) and (2.1.9) are fin<strong>it</strong>e. We will be more precise in chapter<br />

five.<br />

W<strong>it</strong>h the help of this short introduction, we are ready to analyze more closely the<br />

solutions of Sturm-Liouville problems.<br />

2.2 Orthogonal functions<br />

Two functions u,v ∈ C 0 ( Ī) are said to be orthogonal when (u,v)w = 0 for some<br />

weight function w.<br />

Let us assume that the function a in (1.1.1) vanishes at the endpoints of the interval<br />

I (if I is not bounded, limx→±∞ a(x) = 0). Then, we have the following fundamental<br />

result.<br />

Theorem 2.2.1 - Let {un} n∈N be a sequence of solutions of (1.1.1) corresponding to<br />

the eigenvalues {λn} n∈N . Let us require that λn = λm if n = m. Then one has<br />

(2.2.1)<br />

<br />

I<br />

unum w dx = 0, ∀n,m ∈ N w<strong>it</strong>h n = m.<br />

Proof - Let us multiply the equation (1.1.1) by a differentiable function v and integrate<br />

in I. Recalling the assumptions on a, after integration by parts, one gets<br />

<br />

(2.2.2)<br />

au ′ v ′ <br />

dx + buv dx = λ uvw dx.<br />

I<br />

I<br />

Let n = m, thus relation (2.2.2) is satisfied e<strong>it</strong>her when u = un, v = um, λ = λn, or<br />

when u = um, v = un, λ = λm. In both cases the left-hand sides coincide. Having<br />

λn = λm, this implies (2.2.1).<br />

I


24 Polynomial Approximation of Differential Equations<br />

In particular, theorem 2.2.1 shows that all the families of polynomials introduced in<br />

chapter one are orthogonal w<strong>it</strong>h respect to the inner product (2.1.8), where w is their<br />

corresponding weight function. More precisely, for any n,m ∈ N, n = m, one has<br />

(2.2.3) (Jacobi)<br />

(2.2.4) (Legendre)<br />

(2.2.5) (Chebyshev)<br />

(2.2.6) (Laguerre)<br />

(2.2.7) (Herm<strong>it</strong>e)<br />

1<br />

−1<br />

P (α,β)<br />

n (x)P (α,β)<br />

m (x) (1 − x) α (1 + x) β dx = 0,<br />

+∞<br />

0<br />

1<br />

1<br />

−1<br />

+∞<br />

−∞<br />

−1<br />

Pn(x)Pm(x) dx = 0,<br />

Tn(x)Tm(x)<br />

dx<br />

√ 1 − x 2<br />

= 0,<br />

L (α)<br />

n (x)L (α)<br />

m (x) x α e −x dx = 0,<br />

Hn(x)Hm(x) e −x2<br />

dx = 0.<br />

Note that scaled Laguerre (or Herm<strong>it</strong>e) functions are not orthogonal.<br />

Moreover, considering that b ≡ 0, (2.2.1) and (2.2.2) imply that the orthogonal polyno-<br />

mials also satisfy<br />

(2.2.8)<br />

<br />

I<br />

au ′ nu ′ m dx = 0, ∀n,m ∈ N w<strong>it</strong>h n = m.<br />

This shows that the derivatives are orthogonal polynomials w<strong>it</strong>h respect to the weight<br />

function a.<br />

For any n ≥ 1, if p is a polynomial of degree at most n − 1 we have <br />

since p is a linear combination of the uk’s for k ≤ n − 1.<br />

I punw dx = 0,<br />

By taking x = cos θ in (2.2.5) and recalling (1.5.6), one obtains the well-known<br />

orthogonal<strong>it</strong>y relation for trigonometric functions (see for instance zygmund(1988)):<br />

(2.2.9)<br />

π<br />

0<br />

cos nθ cos mθ dθ = 0, ∀n,m ∈ N, n = m.<br />

This will allows us further characterize Chebyshev polynomials.


Orthogonal<strong>it</strong>y 25<br />

As we promised, we give the proof of theorem 1.1.1.<br />

Proof - For n = 2 we have to determine three degrees of freedom ρ2, σ2, τ2, in order<br />

to evaluate a second degree polynomial. Therefore, (1.1.2) is trivially satisfied. If n > 2,<br />

we wr<strong>it</strong>e p := un − ρnxun−1, where ρn is such that p is a polynomial whose degree is<br />

at most n − 1. Thus, due to the orthogonal<strong>it</strong>y cond<strong>it</strong>ion<br />

<br />

I<br />

pumw dx =<br />

<br />

I<br />

unumw dx − ρn<br />

<br />

I<br />

un−1(xum)w dx = 0, ∀m < n − 2<br />

(note that the degree of xum is less than n − 1). This implies that p is only a linear<br />

combination of the terms un−1 and un−2, i.e., p = σnun−1 + τnun−2. The proof is<br />

concluded.<br />

Finally, norms can be also evaluated.<br />

Theorem 2.2.2 - For any n ∈ N, we have<br />

(2.2.10) (Jacobi)<br />

(2.2.11) (Legendre)<br />

(2.2.12) (Chebyshev)<br />

=<br />

⎧<br />

⎨<br />

⎩<br />

1<br />

−1<br />

<br />

P (α,β)<br />

n<br />

2 (x)<br />

2 α+β+1 Γ(α+1) Γ(β+1)<br />

Γ(α+β+2)<br />

2 α+β+1<br />

(2n+α+β+1) n!<br />

1<br />

−1<br />

1<br />

−1<br />

T 2 n(x)<br />

(1 − x) α (1 + x) β dx<br />

Γ(n+α+1) Γ(n+β+1)<br />

Γ(n+α+β+1)<br />

P 2 n(x) dx =<br />

dx<br />

√ 1 − x 2 =<br />

2<br />

2n + 1 ,<br />

if n = 0,<br />

if n > 0,<br />

⎧<br />

⎨π<br />

if n = 0,<br />

⎩<br />

π<br />

2 if n > 0,


26 Polynomial Approximation of Differential Equations<br />

(2.2.13) (Laguerre)<br />

(2.2.14) (Herm<strong>it</strong>e)<br />

+∞<br />

0<br />

<br />

L (α)<br />

2 n (x) x α e −x dx =<br />

+∞<br />

−∞<br />

H 2 n(x) e −x2<br />

Proof - Taking u = v = un in (2.2.2), if b ≡ 0 we have<br />

(2.2.15) un 2 w = λ −1<br />

n u ′ n 2 a.<br />

Γ(n + α + 1)<br />

,<br />

n!<br />

dx = 2 n n! √ π.<br />

Let us consider first the Jacobi case. We get λn = n(n + α + β + 1) by theorem<br />

1.3.1. Moreover, recalling (1.3.6) and subst<strong>it</strong>uting in (2.2.15), by <strong>it</strong>erating n times this<br />

procedure one obtains<br />

(2.2.16)<br />

= (n + α + β + 1) · · · (n + α + β + n)<br />

= n + α + β + 1<br />

1<br />

−1<br />

<br />

4n<br />

P (α,β)<br />

n<br />

1<br />

−1<br />

2 (x)<br />

<br />

(1 − x) α (1 + x) β dx<br />

P (α+1,β+1)<br />

n−1<br />

4 n n!<br />

=<br />

1<br />

−1<br />

2 (x)<br />

Γ(2n + α + β + 1)<br />

2 2n n! Γ(n + α + β + 1)<br />

<br />

(1 − x) α+1 (1 + x) β+1 dx = · · · =<br />

P (α+n,β+n)<br />

0<br />

2 (x) (1 − x) α+n (1 + x) β+n dx<br />

1<br />

(1 − x)<br />

−1<br />

α+n (1 + x) β+n dx.<br />

We arrive at (2.2.10) by using (1.2.3) w<strong>it</strong>h y = α + n + 1 and x = β + n + 1. Then,<br />

setting α = β = 0 we have (2.2.11). By choosing α = β = −1/2 and considering<br />

(1.5.1) we have (2.2.12) (otherwise recall (1.5.6) and take x = cos θ). To prove (2.2.13)<br />

and (2.2.14) we argue exactly as we did for the Jacobi case. This time we use the<br />

relations (1.6.6) and (1.7.8) respectively.


Orthogonal<strong>it</strong>y 27<br />

Let us now define<br />

(2.2.17) ũn := γ −1<br />

n un, where γn := lim<br />

x→+∞<br />

un(x)<br />

.<br />

xn Of course, we get ũn(x) = x n +{lower degree terms}. The values of γn can be deduced in<br />

the different cases from (1.3.5), (1.5.5), (1.6.4) and (1.7.5). Then, we have the following<br />

minimizing property.<br />

Theorem 2.2.3 - For any n ∈ N and for any polynomial p of degree n such that<br />

p(x) = x n + {lower degree terms}, we have<br />

(2.2.18) ũnw ≤ pw.<br />

Proof - Let us note that ũn − p has degree at most n − 1. Consequently, ũn − p is<br />

orthogonal to ũn. Hence<br />

(2.2.19) ũn 2 w = (ũn,p)w.<br />

Applying the Schwarz inequal<strong>it</strong>y (2.1.7) to the right-hand side of (2.2.19) we get (2.2.18).<br />

2.3 Fourier coefficients<br />

Hereafter, Pn denotes the linear space of polynomials whose degree is at most n. The<br />

dimension of Pn is clearly n + 1. The fact that the polynomials uk, 0 ≤ k ≤ n,<br />

are orthogonal w<strong>it</strong>h respect to some inner product implies that they form a basis for<br />

Pn. Therefore, for any p ∈ Pn, one can determine in a unique way n + 1 coefficients<br />

ck, 0 ≤ k ≤ n, such that<br />

(2.3.1) p =<br />

n<br />

k=0<br />

ckuk.<br />

The ck’s are called the Fourier coefficients of p w<strong>it</strong>h respect to the designated basis.<br />

By knowing the coefficients of p, the value of p(x) at a given x can be efficiently evaluated<br />

by (1.1.2) and (2.3.1) w<strong>it</strong>h a computational cost proportional to n.


28 Polynomial Approximation of Differential Equations<br />

If p = n k=0 ckuk and q = n k=0 bkuk<br />

proven:<br />

, then the following relations are trivially<br />

n<br />

(2.3.2) p + q = (ck + bk)uk,<br />

(2.3.3) λp =<br />

(2.3.4) pq =<br />

k=0<br />

n<br />

(λck)uk, ∀λ ∈ R,<br />

k=0<br />

n<br />

k=0 j=0<br />

n<br />

(ckbj)ukuj.<br />

By integrating (2.3.4) in I, orthogonal<strong>it</strong>y implies<br />

n<br />

(2.3.5) (p,q)w =<br />

Hence<br />

(2.3.6) pw =<br />

m=0<br />

n<br />

m=0<br />

cmbm um 2 w.<br />

c 2 m um 2 w<br />

1<br />

2<br />

.<br />

The quant<strong>it</strong>ies um 2 w, m ∈ N, have been computed in theorem 2.2.2 for the different<br />

families of polynomials. Formula (2.3.6) allows us to evaluate the weighted norm of p<br />

when <strong>it</strong>s coefficients are known. In particular, if we take q ≡ uk, 0 ≤ k ≤ n, in (2.3.5)<br />

(then bm = δkm), the following explic<strong>it</strong> expression for the coefficients is available:<br />

(2.3.7) ck = (p,uk)w<br />

uk 2 w<br />

, p ∈ Pn, 0 ≤ k ≤ n.<br />

Given g : R 2 → R , one can try to characterize the Fourier coefficients of g(x,p(x)),<br />

where p is a polynomial in Pn, in terms of the coefficients of p. This turns out to be useful<br />

for the approximation of nonlinear terms in differential equations. For instance, we can<br />

easily handle the case g(x,p(x)) = xp(x). Namely, from the coefficients of the expression<br />

p = n<br />

k=0 ckuk we want to determine those of the expression xp = n+1<br />

k=0 bkuk. First<br />

of all we deduce from (1.1.2)<br />

(2.3.8) xuk = 1<br />

ρk+1<br />

(uk+1 − σk+1uk − τk+1uk−1), 1 ≤ k ≤ n.


Orthogonal<strong>it</strong>y 29<br />

Therefore, one obtains<br />

(2.3.9) bk =<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

cn<br />

ρn+1<br />

cn−1<br />

ρn<br />

ck−1<br />

ρk<br />

− c0σ1<br />

ρ1<br />

− cnσn+1<br />

ρn+1<br />

− ckσk+1<br />

ρk+1<br />

− c1τ2<br />

ρ2<br />

− ck+1τk+2<br />

ρk+2<br />

k = n + 1,<br />

k = n,<br />

1 ≤ k ≤ n − 1,<br />

k = 0.<br />

For the various families of polynomials, the sequences {ρn}, {σn}, {τn} are respectively<br />

given by (1.3.8), (1.4.2), (1.5.2), (1.6.5) and (1.7.6) (ρ1 and σ1 are defined according to<br />

u0 and u1). For ultraspherical and Herm<strong>it</strong>e polynomials we have σn = 0, ∀n ≥ 1.<br />

For more general functions g the answer can be very complicated and more explic<strong>it</strong><br />

expressions are often unavailable. An interesting case is g(x,p(x)) = p 2 (x). For ultra-<br />

spherical polynomials the following relation holds<br />

(2.3.10) P (α,α)<br />

k P (α,α)<br />

j =<br />

ν Γ(k + ν + 1<br />

2<br />

min(k,j) k + j + ν − 2m<br />

×<br />

(m + ν)(k − m + ν)(j − m + ν)<br />

m=0<br />

<br />

k + j − 2m + 2ν<br />

×<br />

k + j − 2m<br />

where ν = α + 1<br />

2 .<br />

−1<br />

1 ) Γ(j + ν + 2 ) Γ(ν + 1)<br />

2 Γ(k + 2ν) Γ(j + 2ν) Γ(ν + 1<br />

2 )<br />

<br />

m + ν k − m + ν j − m + ν<br />

m k − m j − m<br />

Γ(k + j − m + 2ν) Γ(k + j − 2m + 2ν + 1)<br />

Γ(k + j − m + ν + 1)<br />

k+j−2m<br />

Γ(k + j − 2m + ν + 1<br />

(α,α)<br />

P<br />

2 )<br />

<br />

α > −1, α = −1 2 , k,j ∈ N,<br />

,


30 Polynomial Approximation of Differential Equations<br />

If α = −1 2 , by (1.5.6) we get<br />

(2.3.11) TkTj = 1<br />

<br />

2<br />

Tk+j + T |k−j| , k,j ∈ N.<br />

Besides, for Herm<strong>it</strong>e polynomials we have<br />

(2.3.12) HkHj =<br />

min(k,j) <br />

m=0<br />

2 m m!<br />

<br />

k j<br />

Hk+j−2m, k,j ∈ N.<br />

m m<br />

Taking p = q in (2.3.4), we can finally evaluate the expansion coefficients of p 2 in<br />

terms of those of p. The reader can find more results, suggestions and references in<br />

askey(1975), lecture 5.<br />

2.4 The projection operator<br />

More generally, in analogy w<strong>it</strong>h (2.3.7), if f is a given function such that fw is integrable<br />

in I (when I is not bounded fw will be also required to decay to zero at infin<strong>it</strong>y), then<br />

we define <strong>it</strong>s Fourier coefficients by setting<br />

(2.4.1) ck :=<br />

1<br />

<br />

fukw dx, k ∈ N.<br />

uk 2 w<br />

I<br />

Let us introduce now the operator Πw,n : C 0 ( Ī) → Pn, n ∈ N. This operator acts as<br />

follows. For any continuous function f, we compute <strong>it</strong>s coefficients according to (2.4.1).<br />

Then, Πw,nf ∈ Pn is defined to be the polynomial pn := n<br />

k=0 ckuk. It turns out that<br />

Πw,n is a linear operator. We call pn the orthogonal projection of f onto Pn, through<br />

the inner product (·, ·)w. Of course, one has<br />

(2.4.2) Πw,np = p, ∀p ∈ Pn,<br />

(2.4.3) Πw,num = 0, ∀m > n.


Orthogonal<strong>it</strong>y 31<br />

From (2.3.8), we get for instance<br />

(2.4.4) Πw,n(xun) = − σn+1<br />

ρn+1<br />

un − τn+1<br />

un−1, n ≥ 1.<br />

ρn+1<br />

Taking a truncation of the sum in (2.3.10) or (2.3.12) in order that k + j − 2m ≤ n, we<br />

obtain an expression for Πw,n(ukuj), k,j ∈ N. Therefore, by the linear<strong>it</strong>y of Πw,n,<br />

one obtains the coefficients of Πw,np 2 in terms of the coefficients of p.<br />

The main properties of the projection operator are investigated in section 6.2.<br />

There, under su<strong>it</strong>able hypotheses, we shall see that limn→+∞ Πw,nf = f, where the<br />

lim<strong>it</strong>ing process must be correctly interpreted.<br />

2.5 The maximum norm<br />

Let us assume that I is a bounded interval. In the linear space of continuous functions<br />

in Ī, we define the norm<br />

(2.5.1) u∞ := max<br />

x∈ Ī<br />

|u(x)|, ∀u ∈ C 0 ( Ī).<br />

One easily verifies that ·∞ actually satisfies all the properties (2.1.3)-(2.1.5). We call<br />

u∞ the maximum norm of u in Ī. It can be proven that <strong>it</strong> is not possible to define<br />

an inner product that generates the maximum norm through a relation like (2.1.6).<br />

Two celebrated theorems give a bound to the maximum norm of the derivative<br />

of polynomials. The first one is due to A.Markoff and is successively generalized to<br />

higher order derivatives by his brother (see markoff(1889) and markoff(1892)). The<br />

second one is due to S.Bernstein (bernstein(1912a)). Other references can be found in<br />

jackson(1930), cheney(1966) as well in zygmund(1988), Vol.2. Further refinements<br />

are analyzed in section 3.9. Applications are illustrated in chapter six.<br />

Theorem 2.5.1 (Markoff) - Let Ī = [−1,1], then for any n ∈ N we have<br />

(2.5.2) p ′ ∞ ≤ n 2 p∞, ∀p ∈ Pn.


32 Polynomial Approximation of Differential Equations<br />

Theorem 2.5.2 (Bernstein) - Let Ī = [−1,1], then for any n ∈ N we have<br />

(2.5.3) |p ′ (x)| ≤<br />

n<br />

√ 1 − x 2 p∞, ∀x ∈ I, ∀p ∈ Pn.<br />

Recalling defin<strong>it</strong>ion (2.2.17), we finally state one of the main results characterizing<br />

Chebyshev polynomials (compare w<strong>it</strong>h theorem 2.2.3).<br />

Theorem 2.5.3 - For any n ∈ N and for any polynomial p in<br />

such that p(x) = x n + {lower degree terms}, we have<br />

(2.5.4) ˜ Tn∞ ≤ p∞.<br />

The proof is given for instance in rivlin(1974). Note that<br />

(2.5.5) ˜ Tn∞ =<br />

2.6 Basis transformations<br />

1 if n = 0,<br />

2 1−n if n ≥ 1.<br />

Ī = [−1,1] of degree n<br />

An interesting question is how to transform the Fourier coefficients of a given polynomial<br />

corresponding to an assigned orthogonal basis, into the coefficients of another basis<br />

orthogonal w<strong>it</strong>h respect to a different weight function. The goal is to determine the<br />

so-called connection coefficients of the expansion of any element of the first basis in<br />

terms of the elements of the second basis. For a survey of the major results, we suggest<br />

that the interested reader consult lecture 7 from the book of askey(1975). Here we


Orthogonal<strong>it</strong>y 33<br />

only mention the relation existing between some Jacobi polynomials and Chebyshev<br />

polynomials, namely<br />

(2.6.1) P (α,α)<br />

n<br />

×<br />

k=0<br />

=<br />

<br />

Γ(2α + 1) Γ(n + α + 1)<br />

Γ(α + 1) Γ(n + 2α + 1)<br />

n Γ(k + α + 1<br />

2<br />

<br />

1 Γ(α + 2 ) 2<br />

k! (n − k)!<br />

) Γ(n − k + α + 1<br />

2 )<br />

T |n−2k|<br />

<br />

, n ∈ N, α > − 1<br />

2 .<br />

A problem related to the previous one, is the determination of the weighted norm<br />

of a polynomial, when the expansion coefficients, corresponding to a basis orthogonal<br />

w<strong>it</strong>h respect to another weight function, are known. We give an example. Sometimes,<br />

in practical applications <strong>it</strong> is useful to evaluate the integral 1<br />

−1 p2 dx, p ∈ Pn. If<br />

p is known in terms of <strong>it</strong>s Legendre coefficients, (2.3.6) is the formula we are looking<br />

for. The s<strong>it</strong>uation becomes more complicated when p is given in terms of the Fourier<br />

coefficients of another basis, say the Chebyshev basis. Let us assume that<br />

(2.6.2) p =<br />

Then<br />

(2.6.3)<br />

1<br />

−1<br />

p 2 dx =<br />

n<br />

k=0<br />

ckTk.<br />

The help of some trigonometry leads to the formula<br />

(2.6.4) Ikj =<br />

n n<br />

ckcjIkj,<br />

1<br />

where Ikj = TkTj dx.<br />

k=0 j=0<br />

−1<br />

π<br />

0<br />

cos kθ cos jθ sinθ dθ<br />

⎧<br />

⎪⎨<br />

0 if k + j is odd,<br />

=<br />

⎪⎩<br />

1 1<br />

+<br />

1 − (k + j) 2 1 − (k − j) 2 if k + j is even.


34 Polynomial Approximation of Differential Equations<br />

We note that (2.6.3) has a computational cost proportional to n 2 , while the cost of<br />

(2.3.6) is only proportional to n.<br />

Generalizations to other bases can be also considered, but they are outside the aims<br />

of this book.


3<br />

NUMERICAL INTEGRATION<br />

As we shall see in this chapter, orthogonal polynomials allow us to generate in a very<br />

elegant way high order integration formulas, known as Gauss formulas. Since the theory<br />

is based on the knowledge of the zeroes of the polynomials, the first part is devoted to<br />

the characterization of their main properties.<br />

3.1 Zeroes of orthogonal polynomials<br />

It is well-known that a polynomial of degree n has at most n distinct complex zeroes.<br />

In general, very l<strong>it</strong>tle can be said about real zeroes. An amazing theorem gives a precise<br />

and complete answer in the case of orthogonal polynomials.<br />

Theorem 3.1.1 - Let {un} n∈N be a sequence of solutions of (1.1.1), where un is a<br />

polynomial of degree n, satisfying the orthogonal relation (2.2.1). Then, for any n ≥ 1,<br />

un has exactly n real distinct zeroes in I.<br />

Proof - We first note that <br />

I un(x)w(x)dx = 0. Therefore un changes sign in I, hence<br />

<strong>it</strong> has at least one real zero ξ1 ∈ I. If n = 1 the proof is ended. If n > 1, then we can<br />

find another zero ξ2 ∈ I of un, w<strong>it</strong>h ξ2 = ξ1, since if un vanished only at ξ1, then the<br />

polynomial (x − ξ1)un would not change sign in I, which is in contradiction w<strong>it</strong>h the<br />

relation <br />

I (x − ξ1)un(x)w(x)dx = 0, obtained by orthogonal<strong>it</strong>y. In a similar fashion,


36 Polynomial Approximation of Differential Equations<br />

one considers the polynomial (x − ξ1)(x − ξ2)un and, if n > 2, deduces the existence<br />

of a third zero, and so on. The procedure ends when n zeroes are finally obtained.<br />

For any n ≥ 1 we denote the n zeroes of un by ξ (n)<br />

k , 1 ≤ k ≤ n. We assume that these<br />

are in increasing order (although many authors prefer the reverse order). It is obvious<br />

that the polynomial u ′ n has n − 1 real distinct zeroes in I. We denote these zeroes by<br />

η (n)<br />

k , 1 ≤ k ≤ n − 1. One can prove that un and un−1 do not have common zeroes.<br />

Moreover, between any two consecutive zeroes of un−1, there exists one and only one<br />

zero of un.<br />

Many general theorems characterize the zeroes of orthogonal polynomials. It suf-<br />

fices to recall the following statement.<br />

Theorem 3.1.2 - Let {un} n∈N be a sequence of orthogonal polynomials in I. Then,<br />

for any interval [a,b] ⊂ I, a < b, <strong>it</strong> is possible to find m ∈ N such that um has at least<br />

one zero in [a,b].<br />

In other words, this theorem states that the set J = <br />

n≥1<br />

n k=1 {ξ(n)<br />

k<br />

} is dense in Ī.<br />

We are going to review some properties of the zeroes of classical orthogonal poly-<br />

nomials. We refer for instance to szegö(1939) for proofs and further results.<br />

Zeroes in the Jacobi case - We restrict ourselves to the case where − 1<br />

2<br />

− 1<br />

2<br />

1 ≤ β ≤ 2 . Under these cond<strong>it</strong>ions, for n ≥ 1 we have the estimates<br />

(3.1.1) −1 ≤ −cos<br />

≤ α ≤ 1<br />

2 and<br />

k + (α + β − 1)/2<br />

k<br />

π ≤ ξ(n)<br />

k ≤ −cos<br />

π ≤ 1,<br />

n + (α + β + 1)/2 n + (α + β + 1)/2<br />

1 ≤ k ≤ n.<br />

This shows that asymptotically the distance between two consecutive zeroes is propor-<br />

tional to 1/n for points located in the central part of the interval I =] − 1,1[, and<br />

proportional to 1/n 2 for points located near the endpoints of the same interval.


Numerical Integration 37<br />

In the case of ultraspherical polynomials (− 1<br />

2<br />

refined as follows:<br />

(3.1.2) −1 < −cos<br />

(3.1.3) 0 ≤ cos<br />

k + α<br />

2<br />

− 1<br />

4<br />

n + α + 1<br />

2<br />

π ≤ ξ (n)<br />

k<br />

n − k + 1<br />

n + α + 1 π ≤ ξ<br />

2<br />

(n)<br />

k<br />

Of course, if n is odd we have ξ (n)<br />

(n+1)/2 = 0.<br />

≤ −cos<br />

1<br />

≤ α = β ≤ 2 ), the inequal<strong>it</strong>ies can be<br />

k<br />

n + α + 1<br />

2<br />

≤ cos n − k + α<br />

2<br />

n + α + 1<br />

2<br />

π ≤ 0, 1 ≤ k ≤ <br />

n<br />

2 ,<br />

+ 3<br />

4<br />

π < 1,<br />

n + 1 − <br />

n<br />

2 ≤ k ≤ n.<br />

Particularly interesting are the following cases, in which an exact expression of the<br />

zeroes is known:<br />

(3.1.4) ξ (n)<br />

k<br />

(3.1.5) ξ (n)<br />

k<br />

(3.1.6) ξ (n)<br />

k<br />

(3.1.7) ξ (n)<br />

k<br />

where 1 ≤ k ≤ n.<br />

2k − 1<br />

= −cos π if α = β = −1<br />

2n 2 ,<br />

k<br />

1<br />

= −cos π if α = β =<br />

n + 1 2 ,<br />

= −cos<br />

2k<br />

2n + 1<br />

1<br />

π if α = , β = −1<br />

2 2 ,<br />

2k − 1<br />

1<br />

= −cos π if α = −1 , β =<br />

2n + 1 2 2 ,<br />

Zeroes in the Legendre case - Estimates indicating the location of the zeroes are obtained<br />

by setting α = 0 in (3.1.2) and (3.1.3). Figure 3.1.1 shows their distribution in the<br />

interval I =] − 1,1[ for n = 7 and n = 10.


38 Polynomial Approximation of Differential Equations<br />

Figure 3.1.1 - Legendre zeroes for n = 7 and n = 10.<br />

Zeroes in the Chebyshev case - Due to (1.5.1) the zeroes are those of ultraspherical<br />

polynomials w<strong>it</strong>h α = β = −1 2 . Hence, they are exactly given by (3.1.4). On the<br />

other hand, thanks to equal<strong>it</strong>y (1.5.6), they can be also obtained from the roots of the<br />

equation: cos nθ = 0, n ≥ 1, θ ∈]0,π[. In figure 3.1.2 we give the distribution of the<br />

Chebyshev zeroes for n = 7 and n = 10.<br />

Figure 3.1.2 - Chebyshev zeroes for n = 7 and n = 10.<br />

Zeroes in the Laguerre case - According to tricomi(1954), p.153, and gatteschi(1964),<br />

a good approximation of the zeroes of Laguerre polynomials is obtained by the following<br />

procedure. For any n ≥ 1, we first find the roots of the equation<br />

(3.1.8) y (n)<br />

k<br />

− sin y(n)<br />

k<br />

3 n − k + 4<br />

= 2π , 1 ≤ k ≤ n.<br />

2n + α + 1<br />

Then we set ˆy (n)<br />

<br />

k := cos 1<br />

2y(n) 2 k , and define for 1 ≤ k ≤ n,


Numerical Integration 39<br />

(3.1.9) z (n)<br />

k := 2(2n + α + 1)ˆy (n)<br />

k<br />

−<br />

1<br />

6(2n + α + 1)<br />

<br />

5<br />

4(1 − ˆy (n)<br />

−<br />

k<br />

)2<br />

1<br />

1 − ˆy (n)<br />

k<br />

− 1 + 3α 2<br />

Let us permute the values of z (n)<br />

k , 1 ≤ k ≤ n, in such a way they are in increasing order.<br />

After this rearrangement, we denote the new sequence by ˆz (n)<br />

k , 1 ≤ k ≤ n. Finally, we<br />

have<br />

(3.1.10) ξ (n)<br />

k ≈ ˆz (n)<br />

k<br />

, 1 ≤ k ≤ n.<br />

By virtue of theorem 3.1.2, one also gets limn→+∞ ξ (n)<br />

n = +∞, limn→+∞ ξ (n)<br />

1 = 0.<br />

Moreover, the largest zero tends to infin<strong>it</strong>y like 4n, while the smallest zero behaves like<br />

π 2<br />

4n . We plot in figure 3.1.3 the Laguerre zeroes for α = 0 and n = 10 in the interval<br />

]0,40[.<br />

Figure 3.1.3 - Laguerre zeroes for α = 0 and n = 10.<br />

Zeroes in the Herm<strong>it</strong>e case - It is sufficient to recall the relations (1.7.9)<br />

and (1.7.10).<br />

Therefore, when n is even, pos<strong>it</strong>ive zeroes are approximated by<br />

<br />

ˆz (n/2)<br />

k , 1 ≤ k ≤ n<br />

2 ,<br />

where these quant<strong>it</strong>ies are obtained w<strong>it</strong>h the procedure described above w<strong>it</strong>h α = −1 2 .<br />

<br />

When n is odd, pos<strong>it</strong>ive zeroes are approximated by<br />

these terms are evaluated for α = 1<br />

2 . Moreover ξ(n)<br />

(n+1)/2<br />

ˆz ((n−1)/2)<br />

k<br />

, 1 ≤ k ≤ n−1<br />

2 , where<br />

= 0. For all n, the zeroes are<br />

symmetrically distributed around x = 0. A plot in the interval ]−6,6[ is given in figure<br />

3.1.4 for n = 15.<br />

Figure 3.1.4 - Herm<strong>it</strong>e zeroes for n = 15.<br />

.


40 Polynomial Approximation of Differential Equations<br />

In the numerical computations, the following algor<strong>it</strong>hm is generally adopted. For<br />

each zero, an in<strong>it</strong>ial guess is obtained as described above (for the Jacobi case the average<br />

between the upper and lower bound in the estimates (3.1.1) or (3.1.2)-(3.1.3) can be<br />

taken into account). A few <strong>it</strong>erations of a Newton method is then sufficient to determine<br />

accurately the zero. The point values of the polynomial and <strong>it</strong>s derivative can be<br />

computed from (1.1.2) and (1.1.3).<br />

Concerning the case of Laguerre (or Herm<strong>it</strong>e), when n grows, rounding errors occur<br />

in the procedure, due to the sharp oscillations exhib<strong>it</strong>ed by the polynomials. Consider-<br />

able improvements are obtained operating w<strong>it</strong>h scaled Laguerre (or Herm<strong>it</strong>e) functions.<br />

Of course, the values of the zeroes are the same, but the derivatives in the neighborhood<br />

of the zeroes are now moderate.<br />

Similar arguments hold for the determination of the zeroes of the derivatives of<br />

orthogonal polynomials. An in<strong>it</strong>ial approximation can obtained by considering that<br />

each zero η (n)<br />

k , 1 ≤ k ≤ n − 1, of u′ n lies between two consecutive zeroes of un. In<br />

alternative, one can use relations (1.3.6), (1.6.6), (1.7.8).<br />

As <strong>it</strong> will be clear in the following sections, for n ≥ 1, in the Jacobi case <strong>it</strong> is<br />

convenient to set η (n)<br />

0 := −1 and η (n)<br />

n := 1 (though these are not in general zeroes).<br />

Particularly interesting are the s<strong>it</strong>uations where an explic<strong>it</strong> expression is known, i.e.<br />

(3.1.11) η (n)<br />

k<br />

(3.1.12) η (n)<br />

k<br />

(3.1.13) η (n)<br />

k<br />

(3.1.14) η (n)<br />

k<br />

= −cos kπ<br />

n<br />

0 ≤ k ≤ n, if α = β = − 1<br />

2 ,<br />

2k + 1<br />

1<br />

= −cos π 1 ≤ k ≤ n − 1, if α = β =<br />

2n + 2 2 ,<br />

2k + 1<br />

1<br />

= −cos π 1 ≤ k ≤ n, if α = , β = −1<br />

2n + 1 2 2 ,<br />

= −cos<br />

2k<br />

1<br />

π 0 ≤ k ≤ n − 1, if α = −1 , β =<br />

2n + 1 2 2 .


Numerical Integration 41<br />

We give the plots of the zeroes of the derivative of Legendre and Chebyshev poly-<br />

nomials for n = 10, in figure 3.1.5.<br />

Figure 3.1.5 - The points η (10)<br />

k , 0 ≤ k ≤ 10, in the cases of Legendre and Chebyshev.<br />

By virtue of (1.5.6) and (1.5.7), Chebyshev polynomials satisfy for n ≥ 1<br />

(3.1.15) Tn(η (n)<br />

j ) = (−1) n+j , 0 ≤ j ≤ n,<br />

(3.1.16) T ′ n(ξ (n)<br />

j ) =<br />

n (−1)n+j<br />

<br />

1 − [ξ (n)<br />

j ] 2<br />

, 1 ≤ j ≤ n.<br />

Other useful relations are (n ≥ 1, α ≥ −1, β ≥ −1)<br />

(3.1.17)<br />

(3.1.18)<br />

(3.1.19)<br />

d<br />

dx<br />

d<br />

dx<br />

(α,β)<br />

P n (1) =<br />

P (α,β)<br />

n<br />

(n + α + β + 1) Γ(n + α + 1)<br />

,<br />

2 (n − 1)! Γ(α + 2)<br />

(−1) = − (−1)n (n + α + β + 1) Γ(n + β + 1)<br />

,<br />

2 (n − 1)! Γ(β + 2)<br />

d<br />

dx L(α)<br />

n (0) = −<br />

Γ(n + α + 1)<br />

(n − 1)! Γ(α + 2) .<br />

These equal<strong>it</strong>ies are easily obtained from (1.3.1) and (1.6.1).


42 Polynomial Approximation of Differential Equations<br />

3.2 Lagrange polynomials<br />

As remarked in section 2.3, the set {uk}0≤k≤n is a basis in the space Pn of polynomials<br />

of degree at most n. When n + 1 distinct points are given in<br />

Ī, another basis in Pn<br />

is generated in a natural way. This is the basis of Lagrange polynomials w<strong>it</strong>h respect<br />

to the prescribed points. An element of the basis attains the value 1 at a certain point<br />

and vanishes in the remaining n points.<br />

Let us analyze first the Jacobi case. We examine two interesting examples. On<br />

one hand we have the set of the Lagrange polynomials in Pn−1 relative to the n points<br />

ξ (n)<br />

k<br />

(α,β)<br />

, 1 ≤ k ≤ n, i.e., the zeroes of P n . On the other we have the set of Lagrange<br />

polynomials in Pn relative to the n + 1 points η (n)<br />

k , 0 ≤ k ≤ n, i.e., the zeroes of<br />

d (α,β)<br />

dxP n plus the two points −1 and 1.<br />

In the first case, the elements of the basis are denoted by l (n)<br />

j , 1 ≤ j ≤ n. These<br />

polynomials in Pn−1 are uniquely defined by the cond<strong>it</strong>ions<br />

(3.2.1) l (n)<br />

j (ξ (n)<br />

i ) =<br />

<br />

1 if i = j<br />

, 1 ≤ j ≤ n.<br />

0 if i = j<br />

They actually form a basis because any polynomial p ∈ Pn−1 can be wr<strong>it</strong>ten as follows:<br />

(3.2.2) p =<br />

n<br />

j=1<br />

p(ξ (n)<br />

j ) l (n)<br />

j .<br />

Therefore, p is a linear combination of the Lagrange polynomials. Such a combination<br />

is uniquely determined by the coefficients p(ξ (n)<br />

j ), 1 ≤ j ≤ n.<br />

The following expression is easily proven:<br />

(3.2.3) l (n)<br />

j (x) =<br />

n<br />

k=1<br />

k=j<br />

x − ξ (n)<br />

k<br />

ξ (n)<br />

j<br />

− ξ(n)<br />

k<br />

, 1 ≤ j ≤ n, x ∈ [−1,1].<br />

For future applications, <strong>it</strong> is more convenient to consider the alternate expression<br />

⎧<br />

⎪⎨<br />

un(x)<br />

if x = ξ (n)<br />

j ,<br />

(3.2.4) l (n)<br />

j (x) =<br />

⎪⎩<br />

u ′ n(ξ (n)<br />

j ) (x − ξ (n)<br />

j )<br />

1 if x = ξ (n)<br />

j ,


Numerical Integration 43<br />

where 1 ≤ j ≤ n and un = P (α,β)<br />

n . Of course, we have<br />

(3.2.5) lim<br />

x→ξ (n)<br />

j<br />

l (n)<br />

j (x) = lim<br />

x→ξ (n)<br />

j<br />

u ′ n(x)<br />

u ′ n(ξ (n)<br />

j )<br />

= 1, 1 ≤ j ≤ n.<br />

Plots of the Lagrange basis w<strong>it</strong>h respect to the Legendre and Chebyshev zeroes are<br />

given in figures 3.2.1 and 3.2.2 for n = 7.<br />

Figure 3.2.1 - Legendre Lagrange basis Figure 3.2.2 - Chebyshev Lagrange basis<br />

w<strong>it</strong>h respect to ξ (7)<br />

i , 1 ≤ i ≤ 7. w<strong>it</strong>h respect to ξ (7)<br />

i , 1 ≤ i ≤ 7.<br />

The other basis we mentioned above is denoted by { ˜ l (n)<br />

j }0≤j≤n ⊂ Pn, and <strong>it</strong> is<br />

defined by<br />

(3.2.6)<br />

˜ l (n)<br />

j (η (n)<br />

j ) = δij, 0 ≤ i ≤ n, 0 ≤ j ≤ n.<br />

Indeed, for any polynomial p ∈ Pn, one has<br />

(3.2.7) p =<br />

n<br />

j=0<br />

p(η (n)<br />

j ) ˜ l (n)<br />

j .


44 Polynomial Approximation of Differential Equations<br />

In analogy w<strong>it</strong>h (3.2.4), we prove the following propos<strong>it</strong>ion.<br />

Theorem 3.2.1 - For any n ≥ 1, we have<br />

(3.2.8) ˜ l (n)<br />

j (x) =<br />

w<strong>it</strong>h x = η (n)<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

(−1) n (n − 1)! Γ(β + 2)<br />

(n + α + β + 1) Γ(n + β + 1) (x − 1)u′ n(x) if j = 0,<br />

(x 2 − 1)u ′ n(x)<br />

n(n + α + β + 1) un(η (n)<br />

j ) (x − η (n)<br />

j )<br />

if 1 ≤ j ≤ n − 1,<br />

(n − 1)! Γ(α + 2)<br />

(n + α + β + 1) Γ(n + α + 1) (x + 1)u′ n(x) if j = n,<br />

j and un = P (α,β)<br />

n . Besides, we have ˜l (n)<br />

j (η (n)<br />

j ) = 1, 0 ≤ j ≤ n.<br />

Proof - If j = 0, then γ(x − 1)u ′ n, γ ∈ R, is a polynomial in Pn vanishing at<br />

η (n)<br />

k , 1 ≤ k ≤ n. After evaluating in x = −1, the constant γ has to be determined<br />

so that −2γu ′ n(−1) = 1. Finally, the value u ′ n(−1) is found from (3.1.18). A similar<br />

procedure applies when j = n (this time recalling (3.1.17)). If 1 ≤ j ≤ n − 1, for some<br />

γ ∈ R we have<br />

(3.2.9) lim<br />

x→η (n)<br />

j<br />

γ (x2 − 1)u ′ n(x)<br />

x − η (n)<br />

j<br />

= lim<br />

x→η (n)<br />

j<br />

γ[(x 2 − 1)u ′′ n(x) + 2xu ′ n(x)]<br />

= γ([η (n)<br />

j ] 2 − 1)u ′′ n(η (n)<br />

j ) = γn(n + α + β + 1)un(η (n)<br />

j ).<br />

The last equal<strong>it</strong>y in (3.2.9) is a consequence of (1.3.1). The required γ is then obtained<br />

by equating the last expression in (3.2.9) to 1.<br />

Recalling (1.5.1), (1.5.4) and (3.1.15), relation (3.2.8) becomes in the Chebyshev<br />

case (see also gottlieb, hussaini and orszag(1984)):


Numerical Integration 45<br />

(3.2.10)<br />

˜ l (n)<br />

j (x) =<br />

⎧<br />

(−1)<br />

⎪⎨<br />

n<br />

2n2 (x − 1)T ′ n(x) if j = 0,<br />

(−1) j+n<br />

n2 (x2 − 1)T ′ n(x)<br />

if 1 ≤ j ≤ n − 1,<br />

⎪⎩<br />

x − η (n)<br />

j<br />

1<br />

2n 2 (x + 1)T ′ n(x) if j = n.<br />

The values η (n)<br />

j , 0 ≤ j ≤ n, are given in (3.1.11).<br />

Lagrange bases w<strong>it</strong>h respect to zeroes of derivatives of Legendre and Chebyshev poly-<br />

nomials are shown for n = 7, respectively in figures 3.2.3 and 3.2.4.<br />

Figure 3.2.3 - Legendre Lagrange basis Figure 3.2.4 - Chebyshev Lagrange basis<br />

w<strong>it</strong>h respect to η (7)<br />

i , 0 ≤ i ≤ 7. w<strong>it</strong>h respect to η (7)<br />

i , 0 ≤ i ≤ 7.<br />

Application of a similar procedure to the case of Laguerre (or Herm<strong>it</strong>e) polynomials<br />

is straightforward. Again, we denote by {l (n)<br />

j }1≤j≤n the Lagrange basis w<strong>it</strong>h respect to<br />

the Laguerre (respectively Herm<strong>it</strong>e) zeroes. Relations (3.2.2) and (3.2.3) are still valid.<br />

Formula (3.2.4) also holds provided un = L (α)<br />

n (respectively un = Hn).


46 Polynomial Approximation of Differential Equations<br />

In particular, for the Laguerre case, another set of Lagrange polynomials in Pn−1 will<br />

be useful in what follows. We first define η (n)<br />

0 := 0, and then we construct the Lagrange<br />

basis { ˜ l (n)<br />

j }0≤j≤n−1 w<strong>it</strong>h respect to η (n)<br />

i , 0 ≤ i ≤ n − 1. W<strong>it</strong>h a proof similar to that<br />

of theorem 3.2.1 (this time using (1.6.1) and (3.1.19)), one gets, for n ≥ 1,<br />

(3.2.11)<br />

˜ l (n)<br />

j (x) =<br />

where un = L (α)<br />

n and x ∈ [0,+∞[.<br />

3.3 The interpolation operators<br />

⎧<br />

(n − 1)! Γ(α + 2)<br />

⎪⎨<br />

− u<br />

Γ(n + α + 1)<br />

⎪⎩<br />

′ n(x) if j = 0,<br />

−<br />

x u ′ n(x)<br />

if 1 ≤ j ≤ n − 1,<br />

n un(η (n)<br />

j ) (x − η (n)<br />

j )<br />

Following the guideline of section 2.4, we introduce some new operators defined in the<br />

space of continuous functions. They are named interpolation operators. We first begin<br />

by considering the set of polynomials {uk}k∈N , orthogonal w<strong>it</strong>h respect to the weight<br />

w. Next, we choose n ≥ 1 and evaluate the zeroes ξ (n)<br />

k , 1 ≤ k ≤ n, of un. At this point,<br />

we define Iw,n : C 0 (I) → Pn−1 to be the operator mapping a continuous function f<br />

to the unique polynomial pn ∈ Pn−1 satisfying pn(ξ (n)<br />

j ) = f(ξ (n)<br />

j ), 1 ≤ j ≤ n. This<br />

polynomial is the interpolant of f w<strong>it</strong>h respect to the zeroes of un. The operator Iw,n<br />

is linear. Moreover, one can check that<br />

(3.3.1) Iw,np = p, ∀p ∈ Pn−1.<br />

A question arose in sections 2.3 and 2.4 on how to characterize a certain expression<br />

g(x,p(x)) in terms of a given polynomial p. A very simple answer is obtained when p<br />

is determined by the values attained in a set of given points. For example, we have<br />

(3.3.2) [Iw,ng(x,p(x))] x=ξ (n)<br />

j<br />

for any polynomial p ∈ Pn−1.<br />

= g(ξ (n)<br />

j ,p(ξ (n)<br />

j )), 1 ≤ j ≤ n,


Numerical Integration 47<br />

This means that, by knowing the values p(ξ (n)<br />

j ), 1 ≤ j ≤ n, we can for instance recover<br />

the point values of Iw,np 2 . These are p 2 (ξ (n)<br />

j ), 1 ≤ j ≤ n. Hence, the value in x of<br />

the interpolant of p 2 is n<br />

j=1 p2 (ξ (n)<br />

j )l (n)<br />

j (x).<br />

This trivial remark is a key point in the treatment of nonlinear terms in the approxi-<br />

mation of differential equations (see section 9.8).<br />

d (α,β)<br />

Another interpolant operator is defined in relation to the zeroes of dxP n . Thus<br />

for n ≥ 1, let us consider Ĩw,n : C0 ([−1,1]) → Pn to be the operator associating to the<br />

function f, the unique polynomial pn ∈ Pn satisfying pn(η (n)<br />

j ) = f(η (n)<br />

j ), 0 ≤ j ≤ n.<br />

The operator is linear and we get<br />

(3.3.3) Ĩw,np = p, ∀p ∈ Pn.<br />

For Laguerre polynomials Ĩw,n : C0 (n)<br />

([0,+∞[) → Pn−1 is characterized by ( Ĩw,nf)(η j )<br />

= f(η (n)<br />

j ), 0 ≤ j ≤ n − 1, where η (n)<br />

0 = 0.<br />

As already mentioned in section 2.4, the analysis when n tends to +∞, of the<br />

operators introduced above, is one of the primary issues of chapter six.<br />

3.4 Gauss integration formulas<br />

We have reached the crucial subject of this chapter, that is the formulas for numerical<br />

integration. Given a polynomial p ∈ Pn, one is concerned w<strong>it</strong>h finding the value of<br />

<br />

I pwdx. If the Fourier expansion of p = n−1 k=0 ckuk is known, the answer has been<br />

given in chapter two, i.e., <br />

pwdx = c0<br />

I I u0wdx. On the other hand, when p is known<br />

through <strong>it</strong>s values at the zeroes of un, <strong>it</strong> is sufficient to integrate (3.2.2) obtaining<br />

(3.4.1)<br />

where w (n)<br />

j := <br />

I l(n) j<br />

<br />

I<br />

pw dx =<br />

n<br />

j=1<br />

p(ξ (n)<br />

j ) w (n)<br />

j ,<br />

wdx, 1 ≤ j ≤ n, are the weights of the integration formula.


48 Polynomial Approximation of Differential Equations<br />

Expression (3.4.1) is known as Gauss quadrature formula. We will also refer to the ξ (n)<br />

j ’s<br />

as the Gauss integration nodes. Of course, once the weights are evaluated, (3.4.1) holds<br />

for any polynomial p ∈ Pn−1.<br />

Until now, we have not made use of the property that the nodes are related to<br />

zeroes of orthogonal polynomials. This assumption is fundamental in proving the next<br />

remarkable result.<br />

Theorem 3.4.1 - Formula (3.4.1) is true for any polynomial p of degree less or equal<br />

2n − 1.<br />

Proof - We take q := Iw,np. Hence, q ∈ Pn−1 and p − q vanishes at the points<br />

ξ (n)<br />

j , 1 ≤ j ≤ n. Then p − q = unr, where r is a polynomial of degree at most n − 1.<br />

Finally, by orthogonal<strong>it</strong>y and by using the exactness of (3.4.1) for polynomials of degree<br />

n − 1, we get<br />

<br />

I<br />

pw dx =<br />

=<br />

<br />

n<br />

j=1<br />

I<br />

qw dx +<br />

<br />

q(ξ (n)<br />

j ) w (n)<br />

j<br />

I<br />

unrw dx =<br />

=<br />

n<br />

j=1<br />

<br />

I<br />

qw dx<br />

p(ξ (n)<br />

j ) w (n)<br />

j .<br />

We leave as exercise the proof of the theorem’s converse, i.e., if formula (3.4.1) holds<br />

for any p ∈ P2n−1 then the nodes are the zeroes of un. The degree of the integration<br />

formula cannot be improved further. In fact, if one takes p = u 2 n ∈ P2n the right-hand<br />

side of (3.4.1) vanishes. On the contrary, the left-hand side is different from zero.<br />

Explic<strong>it</strong> expressions are known for the quant<strong>it</strong>ies w (n)<br />

j , 1 ≤ j ≤ n, in the various<br />

cases. We summarize the formulas.<br />

Jacobi case - For α > −1, β > −1 and n ≥ 1, we have


Numerical Integration 49<br />

(3.4.2) w (n)<br />

j<br />

(2n + α + β) Γ(n + α) Γ(n + β)<br />

= 2α+β<br />

n! Γ(n + α + β + 1)<br />

×<br />

<br />

P (α,β)<br />

n−1 (ξ(n) j ) d<br />

dx<br />

(α,β)<br />

P n (ξ (n)<br />

−1 j ) , 1 ≤ j ≤ n.<br />

For the proof we argue as follows. By equating the leading coefficients, (1.3.5) yields<br />

(3.4.3)<br />

un(x)<br />

x − ξ (n)<br />

j<br />

= (2n + α + β)(2n + α + β − 1)<br />

2n (n + α + β)<br />

un−1(x) + qj(x), 1 ≤ j ≤ n,<br />

where qj ∈ Pn−2, 1 ≤ j ≤ n, and un = P (α,β)<br />

n . Therefore, recalling (3.4.1) and (3.2.4),<br />

(3.4.4) w (n)<br />

j =<br />

=<br />

=<br />

1<br />

un−1(ξ (n)<br />

j )<br />

1<br />

1<br />

−1<br />

(un−1u ′ n)(ξ (n)<br />

j )<br />

1<br />

un−1(ξ (n)<br />

j )<br />

n<br />

i=1<br />

l (n)<br />

j un−1w dx =<br />

(l (n) (n)<br />

j un−1)(ξ i ) w (n)<br />

i<br />

1<br />

(un−1u ′ n)(ξ (n)<br />

j )<br />

(2n + α + β)(2n + α + β − 1)<br />

2n (n + α + β)<br />

Finally, (3.4.2) is obtained w<strong>it</strong>h the help of (2.2.10).<br />

1<br />

−1<br />

un<br />

x − ξ (n)<br />

j<br />

un−1w dx<br />

un−1 2 w, 1 ≤ j ≤ n.<br />

Legendre case - To obtain an expression for the weights, we simply set α = β = 0 in<br />

(3.4.2). This gives<br />

(3.4.5) w (n)<br />

j<br />

2<br />

=<br />

n<br />

<br />

Pn−1(ξ (n)<br />

j )P ′ n(ξ (n)<br />

−1 j ) , 1 ≤ j ≤ n.<br />

Chebyshev case - This is the most remarkable case. We recall (1.5.1), (3.1.16) and the<br />

relation Tn−1(ξ (n)<br />

j ) = (−1) j+n<br />

(3.4.2), one obtains<br />

(3.4.6) w (n)<br />

j<br />

<br />

1 − [ξ (n)<br />

j ] 2 , 1 ≤ j ≤ n. By taking α = β = − 1<br />

2 in<br />

π<br />

= , 1 ≤ j ≤ n.<br />

n


50 Polynomial Approximation of Differential Equations<br />

The set of the Chebyshev zeroes is the only distribution of Jacobi nodes for which<br />

the corresponding Gaussian weights are all the same. This further sets Chebyshev<br />

polynomials apart from the other families of orthogonal polynomials.<br />

Laguerre case - For α > −1 and n ≥ 1, we have<br />

(3.4.7) w (n)<br />

j<br />

Γ(n + α)<br />

= −<br />

n!<br />

This time the counterpart of (3.4.3) is<br />

(3.4.8)<br />

un(x)<br />

x − ξ (n)<br />

j<br />

<br />

L (α)<br />

n−1 (ξ(n) j ) d<br />

dx L(α) n (ξ (n)<br />

−1 j )<br />

= − 1<br />

n un−1(x) + qj(x), 1 ≤ j ≤ n,<br />

, 1 ≤ j ≤ n.<br />

where qj ∈ Pn−2, 1 ≤ j ≤ n, and un = L (α)<br />

n . The proof proceeds like that of the Jacobi<br />

case after recalling (2.2.13).<br />

Herm<strong>it</strong>e case - We have for n ≥ 1<br />

(3.4.9) w (n)<br />

j<br />

= √ π 2 n+1 n!<br />

<br />

H ′ n(ξ (n)<br />

−2 j ) , 1 ≤ j ≤ n.<br />

For the proof we can argue as in the previous cases. We only note that, by (1.7.8) one<br />

has H ′ n(ξ (n)<br />

j ) = 2nHn−1(ξ (n)<br />

j ), 1 ≤ j ≤ n.<br />

We observe that for all the cases the weights are strictly pos<strong>it</strong>ive. This is clear if<br />

we consider that<br />

w (n)<br />

j<br />

=<br />

n<br />

i=1<br />

[l (n)<br />

j ] 2 (ξ (n)<br />

i ) w (n)<br />

i<br />

=<br />

<br />

I<br />

[l (n)<br />

j ] 2 w dx > 0, 1 ≤ j ≤ n.<br />

Once the zeroes have been computed, the weights can be evaluated using the recurrence<br />

formula (1.1.2) to recover the values of un−1 and u ′ n.<br />

Due to their high accuracy, Gauss formulas play a fundamental role in the theoret-<br />

ical analysis of spectral methods. The possibil<strong>it</strong>y of integrating polynomials of degree<br />

2n − 1 just by knowing their values at n points will be widely used. The reader inter-<br />

ested in collecting more results may consult szegö(1939), stroud and secrest(1966),<br />

ghizzetti and ossicini(1970), davis and rabinow<strong>it</strong>z(1984), where other explic<strong>it</strong> ex-<br />

pressions of the weights are also given.


Numerical Integration 51<br />

3.5 Gauss-Lobatto integration formulas<br />

In this section, we study an integration formula based on the nodes obtained from<br />

the derivatives of Jacobi polynomials. Integrating (3.2.6), we get for n ≥ 1 and any<br />

polynomial p ∈ Pn, the so called Gauss-Lobatto formula<br />

(3.5.1)<br />

1<br />

−1<br />

pw dx =<br />

n<br />

j=0<br />

p(η (n)<br />

j ) ˜w (n)<br />

j ,<br />

where w is the Jacobi weight function and ˜w (n)<br />

j := 1<br />

Gauss-Lobatto weights.<br />

−1 ˜l (n)<br />

j<br />

wdx, 0 ≤ j ≤ n, are the<br />

The equation (3.5.1) represents a high degree integration formula by virtue of the fol-<br />

lowing theorem.<br />

Theorem 3.5.1 - Formula (3.5.1) is true for any polynomial p ∈ P2n−1.<br />

Proof - Here we define q := Ĩw,np. Thus p − q = (1 − x 2 )u ′ nr, where r ∈ Pn−2.<br />

Moreover, r is the derivative of some polynomial s ∈ Pn−1. Then, recalling (2.2.8) and<br />

that a = (1 − x 2 )w, one finally obtains<br />

1<br />

−1<br />

pw dx =<br />

=<br />

1<br />

n<br />

j=0<br />

−1<br />

qw dx +<br />

q(η (n)<br />

j ) ˜w (n)<br />

j<br />

1<br />

−1<br />

=<br />

au ′ ns ′ dx =<br />

n<br />

j=0<br />

Formula (3.5.1) does not hold in general when p ∈ P2n.<br />

1<br />

−1<br />

p(η (n)<br />

j ) ˜w (n)<br />

j .<br />

qw dx<br />

One is now left w<strong>it</strong>h the determination of the weights. To this end we prove the following<br />

propos<strong>it</strong>ion.


52 Polynomial Approximation of Differential Equations<br />

Theorem 3.5.2 - For any n ≥ 2, we have<br />

(3.5.2)<br />

˜w (n)<br />

j =<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

2 α+β (β+1) Γ 2 (β+1)(2n+α+β)(n−2)! Γ(n+α)<br />

Γ(n+α+β+2) Γ(n+β+1)<br />

−2α+β <br />

(2n+α+β)Γ(n+α)Γ(n+β)<br />

(n+α+β+1) n! Γ(n+α+β)<br />

P (α,β)<br />

n<br />

2 α+β (α+1) Γ 2 (α+1)(2n+α+β)(n−2)! Γ(n+β)<br />

Γ(n+α+β+2) Γ(n+α+1)<br />

n−1 <br />

(1 − η (n)<br />

m ) if j = 0,<br />

m=1<br />

d (α,β)<br />

dxP n−1<br />

<br />

(η (n)<br />

−1 j )<br />

if 1 ≤ j ≤ n − 1,<br />

n−1 <br />

(1 + η (n)<br />

m ) if j = n.<br />

m=1<br />

Proof - As in section 3.4 for the Gauss weights, we start w<strong>it</strong>h the relation<br />

(3.5.3)<br />

u ′ n(x)<br />

x − η (n)<br />

j<br />

= (2n + α + β)(2n + α + β − 1)<br />

2(n − 1)(n + α + β)<br />

u ′ n−1(x) + q ′ j(x), 1 ≤ j ≤ n − 1,<br />

where qj ∈ Pn−2 and un = P (α,β)<br />

n . To check (3.5.3), <strong>it</strong> is sufficient to compare the highest<br />

degree terms and recall (1.3.5). Thus, if 1 ≤ j ≤ n − 1, by (3.2.8) and (2.2.15) we get<br />

(3.5.4)<br />

˜w (n)<br />

j<br />

=<br />

= −<br />

1<br />

u ′ n−1 (η(n)<br />

j )<br />

=<br />

n<br />

i=0<br />

( ˜l (n)<br />

j u′ n−1)(η (n)<br />

i ) ˜w (n)<br />

i<br />

−1<br />

<br />

n(n + α + β + 1) (u ′ <br />

(n)<br />

n−1un)(η j )<br />

=<br />

1<br />

−1<br />

1<br />

u ′ n−1 (η(n)<br />

j )<br />

a<br />

u ′ n<br />

x − η (n)<br />

j<br />

1<br />

−1<br />

˜ l (n)<br />

j u′ n−1 wdx<br />

u ′ n−1 dx<br />

(2n + α + β)(2n + α + β − 1)<br />

<br />

2n(n − 1)(n + α + β)(n + α + β + 1) (u ′ u<br />

(n)<br />

n−1un)(η j )<br />

′ n−1 2 a<br />

= −<br />

(2n + α + β)(2n + α + β − 1)<br />

<br />

2n(n + α + β + 1) (u ′ (n)<br />

n−1un)(η j )<br />

un−1 2 w.


Numerical Integration 53<br />

The proof is completed using (2.2.10). For the weights relative to the endpoints, we<br />

wr<strong>it</strong>e<br />

(3.5.5) 0 =<br />

1<br />

−1<br />

= 2un(−1)u ′ n−1(−1) ˜w (n)<br />

0<br />

unu ′ n−1(1 − x)w dx =<br />

n<br />

m=0<br />

(unu ′ n−1)(η (n)<br />

m )(1 − η (n)<br />

m ) ˜w (n)<br />

m<br />

− 2α+β (2n + α + β)Γ(n + α)Γ(n + β)<br />

(n + α + β + 1) n! Γ(n + α + β)<br />

n−1 <br />

m=1<br />

(1 − η (n)<br />

m ).<br />

The first equal<strong>it</strong>y in (3.5.5) is due to orthogonal<strong>it</strong>y, while in the last equal<strong>it</strong>y we used<br />

the expression of the weights obtained above for 1 ≤ m ≤ n − 1. Then, by (1.3.3) and<br />

(3.1.18) we deduce the value of ˜w (n)<br />

0 . In the same way one gets the value of ˜w (n)<br />

n .<br />

In the ultraspherical case, since the polynomials are even or odd, the nodes are<br />

symmetrically distributed in [−1,1]. Therefore, when α = β we have n−1<br />

j=m η(n)<br />

m = 0.<br />

Hence, in the Legendre case we obtain<br />

(3.5.6) ˜w (n)<br />

j<br />

=<br />

⎧<br />

2<br />

⎪⎨ n(n + 1)<br />

⎪⎩<br />

−2<br />

<br />

Pn(η<br />

n + 1<br />

(n)<br />

j )P ′ n−1(η (n)<br />

−1 j )<br />

if j = 0 or j = n,<br />

if 1 ≤ j ≤ n − 1.<br />

For the Chebyshev case, relations (3.5.2) are drastically simplified and give<br />

(3.5.7) ˜w (n)<br />

j<br />

=<br />

⎧<br />

π<br />

⎪⎨ 2n<br />

⎪⎩<br />

π<br />

n<br />

if j = 0 or j = n,<br />

if 1 ≤ j ≤ n − 1.<br />

This is a consequence of (3.1.15) and the relation T ′ n−1(η (n)<br />

j ) = −(n − 1)(−1) j+n ,<br />

1 ≤ j ≤ n − 1.<br />

Also for Gauss-Lobatto formulas <strong>it</strong> is possible to prove that the weights are pos<strong>it</strong>ive.<br />

More about the subject can be found in engels(1980) or davis and rabinow<strong>it</strong>z(1984).<br />

Integration formulas of Gauss-Lobatto type allow the determination of the weighted


54 Polynomial Approximation of Differential Equations<br />

integral of polynomials of degree 2n−1, knowing their values only at n+1 points. W<strong>it</strong>h<br />

respect to Gauss formulas there is a slight loss in accuracy. Nevertheless, the fact that<br />

the points x = −1 and x = 1 are included in the nodes is important when imposing<br />

boundary cond<strong>it</strong>ions in the approximation of differential equations (see section 7.4).<br />

3.6 Gauss-Radau integration formulas<br />

Other integration formulas, known as Gauss-Radau formulas, are based on the nodes<br />

given in (3.1.12), (3.1.13) and (3.1.14) respectively. The first case actually corresponds<br />

to a Gauss type integration formula since, thanks to (1.3.6), the nodes given in (3.1.12)<br />

are the zeroes of P (3/2,3/2)<br />

n−1 . Therefore, the relative formula is exact for polynomials up<br />

to degree 2n − 3. The other two sets of n nodes include the points x = 1 or x = −1<br />

respectively. Appropriate weights can be defined so that the associated integration<br />

formulas are valid for polynomials of degree up to 2n − 2. Some details are discussed<br />

for instance in ralston(1965) or davis and rabinow<strong>it</strong>z(1984).<br />

In this section we are concerned w<strong>it</strong>h an add<strong>it</strong>ional formula, based on the zeroes of<br />

d<br />

dx L(α)<br />

n plus the point x = 0 (the only boundary point of the interval [0,+∞[). This is<br />

(3.6.1)<br />

+∞<br />

0<br />

pw dx =<br />

n−1 <br />

j=0<br />

p(η (n)<br />

j ) ˜w (n)<br />

j .<br />

˜(n) l j wdx, where the Lagrange<br />

Here w is the Laguerre weight function and ˜w (n)<br />

j := +∞<br />

0<br />

polynomials are defined in (3.2.11). W<strong>it</strong>h a proof similar to that of theorem 3.5.1,<br />

equation (3.6.1) turns out to be true for any polynomial p of degree at most 2n −2. For<br />

the weights, we obtain<br />

(3.6.2) ˜w (n)<br />

j =<br />

⎧<br />

(α + 1) Γ<br />

⎪⎨<br />

⎪⎩<br />

2 (α + 1) (n − 1)!<br />

Γ(n + α + 1)<br />

Γ(n + α)<br />

n!<br />

<br />

L (α)<br />

n (η (n)<br />

j ) d<br />

dx L(α)<br />

n−1 (η(n)<br />

j )<br />

−1<br />

if j = 0,<br />

if 1 ≤ j ≤ n − 1.<br />

These formulas are proven w<strong>it</strong>h an argument similar to that used in theorem 3.5.2.


Numerical Integration 55<br />

3.7 Clenshaw-Curtis integration formulas<br />

The formulas analyzed in the previous sections are very accurate when the weighted<br />

integral of a polynomial is needed. The weight function is strictly related to the set of<br />

nodes in which the polynomial is known. Nevertheless, very often in applications, one<br />

is concerned w<strong>it</strong>h evaluating the integral corresponding to a different weight function.<br />

For example, we would like to compute 1<br />

pdx, where p is determined at some Jacobi<br />

−1<br />

Gauss-Lobatto nodes. This time, after integrating (3.2.7), we end up w<strong>it</strong>h the formula<br />

(3.7.1)<br />

1<br />

−1<br />

p dx =<br />

n<br />

j=0<br />

p(η (n)<br />

j ) χ (n)<br />

j ,<br />

where χ (n)<br />

j := 1<br />

−1 ˜l (n)<br />

j dx, 0 ≤ j ≤ n.<br />

Of course, relation (3.7.1) is exact for all polynomials in the space Pn. Unfortunately,<br />

if the nodes are not related to zeroes of derivatives of Legendre polynomials (in this<br />

case we would obtain: χ (n)<br />

j<br />

= ˜w(n)<br />

j , 0 ≤ j ≤ n), we cannot expect the formula to be<br />

exact for polynomials of degree higher than n. On the other hand, (3.7.1) is possibly<br />

interesting for future applications. For this reason, we give the expression of the weights<br />

in the case of Chebyshev nodes (which are presented in (3.1.11)), when n ≥ 4 is even.<br />

In practice, we have<br />

(3.7.2) χ (n)<br />

j =<br />

⎧<br />

⎪⎩<br />

1<br />

n 2 − 1<br />

if j = 0,<br />

⎪⎨<br />

⎡<br />

2<br />

⎣1 −<br />

n<br />

(−1)j<br />

n2 − 1 +<br />

n/2−1 <br />

<br />

2 2kjπ<br />

cos<br />

1 − 4k2 n<br />

k=1<br />

⎤<br />

⎦ if 1 ≤ j ≤ n<br />

2 ,<br />

χ (n)<br />

n−j<br />

if n<br />

2<br />

+ 1 ≤ j ≤ n.<br />

W<strong>it</strong>h this choice of nodes and weights, relation (3.7.1) is known as Clenshaw-Curtis<br />

formula (see davis and rabinow<strong>it</strong>z(1984)). This is exact for polynomials of degree<br />

not larger than n.


56 Polynomial Approximation of Differential Equations<br />

3.8 Discrete norms<br />

Two polynomials p,q ∈ Pn−1 are uniquely characterized by the values they take at the<br />

Gauss nodes ξ (n)<br />

j , 1 ≤ j ≤ n. Therefore, since pq is a polynomial of degree at most<br />

2n − 2, the inner product <br />

I<br />

pqw dx can be determined w<strong>it</strong>h the help of (3.4.1). The<br />

s<strong>it</strong>uation is different when the polynomials p,q ∈ Pn are distinguished by their values at<br />

the points η (n)<br />

j , 0 ≤ j ≤ n. This time pq ∈ P2n and the Gauss-Lobatto formula (3.5.1)<br />

becomes invalid. This justifies the defin<strong>it</strong>ion of an inner product (·, ·)w,n : Pn×Pn → R,<br />

and a norm · w,n : Pn → R + , given by<br />

(3.8.1) (p,q)w,n :=<br />

(3.8.2) pw,n :=<br />

n<br />

j=0<br />

⎛<br />

⎝<br />

p(η (n)<br />

j ) q(η (n)<br />

j ) ˜w (n)<br />

j , ∀p,q ∈ Pn,<br />

n<br />

j=0<br />

p 2 (η (n)<br />

j ) ˜w (n)<br />

j<br />

⎞<br />

⎠<br />

1<br />

2<br />

, ∀p ∈ Pn.<br />

These expressions are called discrete inner product and discrete norm respectively.<br />

We first compute the discrete norm of un. For the sake of simplic<strong>it</strong>y we restrict ourselves<br />

to the ultraspherical case.<br />

Theorem 3.8.1 - For any n ≥ 1, we have<br />

(3.8.3) un 2 w,n = 2 2α+1 Γ 2 (n + α + 1)<br />

n n! Γ(n + 2α + 1) ,<br />

where un = P (α,α)<br />

n , α > −1.<br />

Proof - The key point is the relation<br />

(3.8.4) u ′ n−1(η (n)<br />

j ) = −<br />

n(n + 2α)<br />

n + α<br />

un(η (n)<br />

j ), 1 ≤ j ≤ n − 1.<br />

To obtain (3.8.4), we refer to szegö(1939), p.71, formula (4.5.7). We subst<strong>it</strong>ute n − 1<br />

for n in the second part of that formula, and then we eliminate un−1 w<strong>it</strong>h the help of


Numerical Integration 57<br />

the first part. Subst<strong>it</strong>uting x = η (n)<br />

j , 1 ≤ j ≤ n − 1, the terms containing u ′ n vanish.<br />

When (3.8.4) is achieved, the proof is straightforward w<strong>it</strong>h the help of (3.5.2), (3.1.17)<br />

and (3.1.18).<br />

In particular, when α = −1 2 , (3.8.4) yields<br />

(3.8.5) Tn 2 w,n = π, ∀n ≥ 1.<br />

For any fixed n ≥ 1, the two norms · w and · w,n defined in Pn are equivalent.<br />

This means that <strong>it</strong> is possible to find two constants γ1 > 0,γ2 > 0, such that<br />

(3.8.6) γ1 pw,n ≤ pw ≤ γ2 pw,n, ∀p ∈ Pn.<br />

This is not surprising, since in a fin<strong>it</strong>e dimensional space (such as Pn) any two norms<br />

are always equivalent. More important is the following statement, which we prove in<br />

the ultraspherical case.<br />

Theorem 3.8.2 - The constants γ1 and γ2 in (3.8.6) do not depend on n.<br />

Proof - Expanding p w<strong>it</strong>h respect to the orthogonal basis {uk}0≤k≤n, we can wr<strong>it</strong>e<br />

p = cnun + r, where r ∈ Pn−1. Using that (3.5.1) is true in P2n−1, one obtains<br />

(3.8.7) p 2 w,n = c 2 n un 2 w,n + 2cn (un,r)w,n + r 2 w,n<br />

= c 2 n un 2 w,n + 2cn(un,r)w + r 2 w = c 2 n<br />

<br />

un 2 w,n − un 2 2<br />

w + pw, ∀p ∈ Pn.<br />

By direct comparison of (2.2.10) and (3.8.3) we get un 2 w,n ≥ un 2 w, ∀n ≥ 1. This<br />

implies γ2 = 1. On the other hand by (2.3.7) and by the Schwarz inequal<strong>it</strong>y (2.1.7), we<br />

get<br />

(3.8.8) c 2 n =<br />

Thus, subst<strong>it</strong>uting in (3.8.7),<br />

(p,un)w<br />

un 2 w<br />

2<br />

≤ p2 w<br />

un2 , ∀p ∈ Pn.<br />

w


58 Polynomial Approximation of Differential Equations<br />

(3.8.9) p 2 w,n ≤ un 2 w,n<br />

un 2 w<br />

p 2 w, ∀p ∈ Pn.<br />

Again, from a direct computation, we have that the ratio on the right-hand side of<br />

(3.8.9) is bounded by max {2,3 + 2α} = γ −2<br />

1 .<br />

It is possible to compute the norm pw of a polynomial p ∈ Pn as a function of<br />

the values <strong>it</strong> attains at the points η (n)<br />

j , 0 ≤ j ≤ n. This is shown by the following<br />

propos<strong>it</strong>ion.<br />

Theorem 3.8.3 - For any n ≥ 1, we have<br />

(3.8.10) p 2 w = p 2 w,n − un 2 w,n − un 2 w<br />

un 4 w,n<br />

where un = P (α,β)<br />

n , α > −1, β > −1.<br />

Proof - By (2.3.7) and the exactness of (3.5.1) in P2n−1, one has<br />

[(p,un)w,n] 2 , ∀p ∈ Pn,<br />

(3.8.11) cn un 2 w = (p,un)w = (p,un)w,n + cn(un 2 w − un 2 w,n), ∀p ∈ Pn.<br />

Recovering cn, one gets the interesting property<br />

(3.8.12) cn = (p,un)w,n<br />

un2 , ∀p ∈ Pn.<br />

w,n<br />

The proof is concluded after subst<strong>it</strong>ution of cn in (3.8.7).<br />

The right-hand side of (3.8.10) actually depends only on the values of p at the<br />

nodes. Using (2.2.10) and (3.8.3), equation (3.8.10) becomes<br />

(3.8.13) pw =<br />

<br />

p 2 w,n −<br />

n(n + 1)<br />

2(2n + 1)<br />

[(p,Pn)w,n] 2<br />

1<br />

2<br />

, ∀p ∈ Pn,


Numerical Integration 59<br />

(3.8.14) pw =<br />

<br />

p 2 w,n − 1<br />

2π<br />

2<br />

[(p,Tn)w,n]<br />

1<br />

2<br />

, ∀p ∈ Pn,<br />

when specialized to the Legendre and Chebyshev cases respectively.<br />

The result of theorem 3.8.3 can be generalized as follows:<br />

(3.8.15) (p,q)w = (p,q)w,n − un 2 w,n − un 2 w<br />

un 4 w,n<br />

We leave the proof as exercise.<br />

(p,un)w,n (q,un)w,n, ∀p,q ∈ Pn.<br />

Another quant<strong>it</strong>y we would like to compute is the integral 1<br />

−1 p2 dx, in terms of<br />

the values of p ∈ Pn at some nodes. When the nodes are the Legendre Gauss-Lobatto<br />

points, we can use (3.8.13). On the contrary, we are in a s<strong>it</strong>uation similar to that of<br />

section 3.7. We only solve the problem for Chebyshev Gauss-Lobatto nodes. Here we<br />

have p 2 ∈ P2n, while (3.7.1) is satisfied only in Pn. The idea is to duplicate the number<br />

of points and still use (3.7.1). Recalling (3.1.11) and (3.1.4), in the Chebyshev case we<br />

have the remarkable result that<br />

(3.8.16) {ξ (n)<br />

i }1≤i≤n ∪ {η (n)<br />

j }0≤j≤n = {η (2n)<br />

k }0≤k≤2n.<br />

Since p is known at the points (3.1.11) and p(η (n)<br />

j ) = p(η (2n)<br />

2j ), 0 ≤ j ≤ n, we can<br />

extrapolate <strong>it</strong>s value at the points (3.1.4) w<strong>it</strong>h the help of (3.2.7):<br />

n<br />

(3.8.17) p(η (2n)<br />

2i−1 ) = p(ξ(n) i ) =<br />

j=0<br />

Now, we can use (3.7.1), obtaining ∀p ∈ Pn<br />

(3.8.18)<br />

=<br />

n<br />

j=0<br />

1<br />

−1<br />

p 2 dx =<br />

p 2 (η (n)<br />

j ) χ (2n)<br />

2j +<br />

2n<br />

m=0<br />

n<br />

<br />

n<br />

i=1<br />

p(η (n)<br />

j ) ˜ l (n)<br />

j (ξ (n)<br />

i ), 1 ≤ i ≤ n.<br />

k=0<br />

p 2 (η (2n)<br />

m ) χ (2n)<br />

m<br />

p(η (n)<br />

k ) ˜ l (n)<br />

k (ξ(n)<br />

i )<br />

2<br />

χ (2n)<br />

2i−1 .<br />

Finally, the weights can be determined from (3.7.2), and the coefficients ˜ l (n)<br />

k (ξ(n)<br />

i )<br />

from (3.2.10) and (3.1.16). This procedure has a computational cost proportional to n 2 ,<br />

versus the cost of implementing (3.8.10), which is proportional to n.


60 Polynomial Approximation of Differential Equations<br />

A discrete inner product w<strong>it</strong>h <strong>it</strong>s relative norm, can also be conveniently defined<br />

for polynomials in Pn−1 distinguished by their values at the n nodes including the n−1<br />

d zeroes of dxL(α) n<br />

and the point x = 0. However, this defin<strong>it</strong>ion would be redundant.<br />

Indeed, when r,s ∈ Pn−1 we have p := rs ∈ P2n−2; therefore, by virtue of (3.6.1),<br />

the new inner product coincides w<strong>it</strong>h the usual one, i.e. · w, where w is the Laguerre<br />

weight function.<br />

3.9 Discrete maximum norms<br />

Discrete equivalents of the maximum norm, introduced in section 2.5, can be defined.<br />

We study two cases. For any n ≥ 1, we first consider ||| · |||w,n : Pn−1 → R + given by<br />

(3.9.1) |||p|||w,n := max<br />

1≤j≤n |p(ξ(n)<br />

j )|, p ∈ Pn−1,<br />

where the ξ (n)<br />

j ’s are the zeroes of P (α,β)<br />

n .<br />

Similarly, for any n ≥ 1, we define ||| · ||| ∗ w,n : Pn → R + as<br />

(3.9.2) |||p||| ∗ w,n := max<br />

0≤j≤n |p(η(n)<br />

j )|, p ∈ Pn,<br />

where the η (n)<br />

j are the zeroes of d (α,β)<br />

dxP n plus the endpoints of the interval<br />

The following two results are easy to derive:<br />

(3.9.3) |||p|||w,n ≤ p∞, ∀p ∈ Pn−1,<br />

(3.9.4) |||p||| ∗ w,n ≤ p∞, ∀p ∈ Pn.<br />

Ī = [−1,1].<br />

Of course, working w<strong>it</strong>hin fin<strong>it</strong>e dimensional spaces, the inequal<strong>it</strong>ies (3.9.3) and (3.9.4)<br />

can be reversed. For instance, ∀n ≥ 1, we can find γ such that


Numerical Integration 61<br />

(3.9.5) p∞ ≤ γ |||p|||w,n, ∀p ∈ Pn−1.<br />

This time, a statement like that of theorem 3.8.2 cannot be proven. In fact, γ ≡ γ(n)<br />

actually grows w<strong>it</strong>h n. Following erdös(1961), <strong>it</strong> is possible to find a constant δ > 0<br />

such that<br />

(3.9.6) γ(n) > 2<br />

π<br />

log(n) − δ.<br />

However, using Chebyshev nodes, we get the relation (see natanson(1965))<br />

(3.9.7) γ(n) < 2<br />

π<br />

log(n) + 1.<br />

Further inequal<strong>it</strong>ies of this type will be covered in section 6.3.<br />

W<strong>it</strong>hin the context of the never ending properties of Chebyshev polynomials, we<br />

state a result which generalizes theorem 2.5.1. The proof is in duffin and schaef-<br />

fer(1941).<br />

Theorem 3.9.1 - Let w(x) = 1/ √ 1 − x 2 , then for any n ≥ 1, we have<br />

(3.9.8) p ′ ∞ ≤ n 2 |||p||| ∗ w,n, ∀p ∈ Pn.<br />

Another theorem, closely related to theorem 2.5.2, is stated in rivlin(1974), p.104.<br />

Theorem 3.9.2 - Let w(x) = 1/ √ 1 − x 2 , then for any n ≥ 1, we have<br />

(3.9.9) |p ′ (ξ (n)<br />

j )| ≤<br />

n<br />

<br />

1 − [ξ (n)<br />

j ] 2<br />

|||p||| ∗ w,n, 1 ≤ j ≤ n, ∀p ∈ Pn.


62 Polynomial Approximation of Differential Equations<br />

3.10 Scaled weights<br />

For large n, the determination of the Laguerre weights can sometimes lead to severe<br />

numerical problems. In fact, for the nodes largest in magn<strong>it</strong>ude, the corresponding<br />

weights are very small and decay qu<strong>it</strong>e fast to zero for increasing values of n. Following<br />

the hints in section 1.6, <strong>it</strong> is advisable to introduce some scaled weights.<br />

First, using (1.6.12) and (1.6.13), we note that, for n ≥ 1,<br />

(3.10.1)<br />

d<br />

dx L(α) n (ξ (n)<br />

j ) =<br />

n + α<br />

n<br />

d<br />

dx ˆ L (α)<br />

n (ξ (n)<br />

j )<br />

n<br />

<br />

k=1<br />

1 + ξ(n) j<br />

4k<br />

Next, starting from (3.4.7), we define a new set of Gauss type weights<br />

(3.10.2) ω (n)<br />

j<br />

= −<br />

Γ(n + α + 1)<br />

4n 2 n!<br />

:= w (n)<br />

j<br />

(4n + ξ (n)<br />

j )<br />

<br />

S (α)<br />

n (ξ (n)<br />

−2 j )<br />

<br />

, 1 ≤ j ≤ n.<br />

<br />

ˆL (α)<br />

n−1 (ξ(n) j ) d<br />

dx ˆ L (α)<br />

n (ξ (n)<br />

−1 j ) , 1 ≤ j ≤ n.<br />

The values of the scaled Laguerre functions at the nodes are evaluated by (1.6.14) and<br />

(1.6.15). Recalling (3.4.1), we obtain the formula<br />

(3.10.3)<br />

where ˆp := pS (α)<br />

n<br />

+∞<br />

p<br />

0<br />

2 (x) x α e −x dx =<br />

n<br />

j=1<br />

ˆp 2 (ξ (n)<br />

j ) ω (n)<br />

j , ∀p ∈ Pn−1,<br />

are used in place of p in numerical computations. The advan-<br />

tages to this approach consist in a better numerical treatment. More details about the<br />

implementation are examined later in section 7.5.<br />

A similar defin<strong>it</strong>ion is given for Gauss-Radau type weights (see (3.6.2)), i.e.,<br />

(3.10.4) ˜ω (n)<br />

0 := ˜w (n)<br />

0<br />

<br />

S (α)<br />

−2 n (0)<br />

= (α + 1) Γ(n + α + 1)<br />

,<br />

n n!


Numerical Integration 63<br />

(3.10.5) ˜ω (n)<br />

j<br />

= Γ(n + α + 1)<br />

4n 2 n!<br />

4n + η (n)<br />

j<br />

ˆL (α)<br />

n (η (n)<br />

j )<br />

where we noted that, for n ≥ 2<br />

(3.10.6)<br />

Again we have<br />

(3.10.7)<br />

where ˆp := pS (α)<br />

n .<br />

d<br />

<br />

S<br />

dx<br />

(α)<br />

n−1 (x)<br />

−1 <br />

:= ˜w (n)<br />

j<br />

<br />

S (α)<br />

n (η (n)<br />

−2 j )<br />

d<br />

dx ˆ L (α)<br />

n−1 (η(n) j ) + ˆ L (α)<br />

n−1 (η(n)<br />

n−1 <br />

j )<br />

m=1<br />

=<br />

+∞<br />

p<br />

0<br />

2 (x) x α e −x dx =<br />

<br />

S (α)<br />

n−1 (x)<br />

−1 n−1 <br />

n−1 <br />

j=0<br />

m=1<br />

1<br />

4m + η (n)<br />

j<br />

1<br />

, x ∈ [0,+∞[.<br />

4m + x<br />

ˆp 2 (η (n)<br />

j ) ˜ω (n)<br />

j , ∀p ∈ Pn−1,<br />

The same procedure applied to the Herm<strong>it</strong>e weights in (3.4.9), yields<br />

(3.10.8) ω (n)<br />

j<br />

=<br />

√ π n!<br />

2 n−1 [(n/2)!] 2<br />

(3.10.9) ω (n)<br />

j<br />

=<br />

:= w (n)<br />

j<br />

<br />

S (−1/2)<br />

n/2 ([ξ (n)<br />

j ] 2 −2 )<br />

<br />

ˆH ′<br />

n(ξ (n)<br />

−2 j ) , 1 ≤ j ≤ n, if n is even,<br />

:= w (n)<br />

j<br />

√ π n!<br />

2 n−1 [((n − 1)/2)!] 2<br />

<br />

S (1/2)<br />

(n−1)/2 ([ξ(n) j ] 2 −2 )<br />

<br />

ˆH ′<br />

n(ξ (n)<br />

−2 j ) , 1 ≤ j ≤ n, if n is odd.<br />

−1<br />

1 ≤ j ≤ n − 1,<br />

Here the ξ (n)<br />

j ’s are the Herm<strong>it</strong>e zeroes. Relations (1.7.9), (1.7.10), (1.7.11) and (1.7.12)<br />

have been taken into account.<br />

,


4<br />

TRANSFORMS<br />

The aim of the previous chapters was to show that a polynomial adm<strong>it</strong>s a double<br />

representation. One is given by <strong>it</strong>s Fourier coefficients w<strong>it</strong>h respect to set of orthogonal<br />

basis functions, the other by the values at the nodes associated w<strong>it</strong>h a su<strong>it</strong>able high<br />

accuracy integration formula. Clearly, <strong>it</strong> is possible to sw<strong>it</strong>ch from one representation<br />

to the other. The way to do <strong>it</strong> is the subject of this chapter.<br />

4.1 Fourier transforms<br />

One of the main results of chapter two was to show that the polynomials introduced in<br />

chapter one were orthogonal in relation to a su<strong>it</strong>able weight function. In add<strong>it</strong>ion, other<br />

polynomials, i.e., the Lagrange polynomials w<strong>it</strong>h respect to a certain set of nodes, have<br />

been presented in chapter three. These too are in general orthogonal. Actually, due to<br />

(3.4.1) and theorem 3.4.1, we have for any fixed n ≥ 1,<br />

(4.1.1)<br />

<br />

I<br />

l (n)<br />

i l (n)<br />

j<br />

wdx =<br />

n<br />

k=1<br />

l (n)<br />

i (ξ (n)<br />

k<br />

j (ξ (n)<br />

k ) w(n)<br />

k =<br />

) l(n)<br />

⎧<br />

⎨0<br />

if i = j,<br />

⎩<br />

w (n)<br />

i<br />

if i = j.<br />

A similar relation holds when Gauss-Radau nodes are considered. When Gauss-Lobatto<br />

points are used, we loose orthogonal<strong>it</strong>y. However, in terms of the discrete inner product<br />

introduced in section 3.8, we still get orthogonal<strong>it</strong>y. In fact, <strong>it</strong> is trivial to show that,<br />

for any n ≥ 1:


66 Polynomial Approximation of Differential Equations<br />

(4.1.2) ( ˜ l (n)<br />

i , ˜ l (n)<br />

j )w,n = δij ˜w (n)<br />

i , 0 ≤ i ≤ n, 0 ≤ j ≤ n.<br />

This is not true anymore when using the inner product given by (2.1.8). In fact, using<br />

(3.8.15), one obtains<br />

(4.1.3)<br />

1<br />

−1<br />

˜ l (n)<br />

i ˜ l (n)<br />

j wdx = δij ˜w (n)<br />

i − un 2 w,n − un 2 w<br />

un 4 w,n<br />

where un = P (α,β)<br />

n , α > −1, β > −1.<br />

un(η (n)<br />

i )un(η (n)<br />

j ) ˜w (n)<br />

i ˜w (n)<br />

j ,<br />

0 ≤ i ≤ n, 0 ≤ j ≤ n,<br />

In general, the last term on the right-hand side of (4.1.3) tends to zero (when n diverges)<br />

faster than the first term, so that the polynomials considered are almost orthogonal.<br />

Thus, for a fixed n ≥ 1, two orthogonal bases are available in Pn−1, i.e., {uk}0≤k≤n−1<br />

and {l (n)<br />

j }1≤j≤n. Therefore, we can define a linear transformation Kn : Pn−1 → Pn−1,<br />

mapping the vector at the point values of a polynomial p into the vector of <strong>it</strong>s Fourier<br />

coefficients. This application is called discrete Fourier transform, and is represented by<br />

a n × n matrix. Each Fourier coefficient is explic<strong>it</strong>ly determined recalling (2.3.7) and<br />

(3.4.1):<br />

(4.1.4) ci =<br />

1<br />

ui 2 w<br />

n<br />

j=1<br />

Therefore, we have the matrix relation<br />

p(ξ (n)<br />

j ) ui(ξ (n)<br />

j ) w (n)<br />

j , 0 ≤ i ≤ n − 1.<br />

(4.1.5) Kn := {kij} 0≤i≤n−1 , where kij :=<br />

0≤j≤n−1<br />

(n)<br />

ui(ξ j+1 ) w(n) j+1<br />

ui2 w<br />

Obviously Kn is invertible. We express <strong>it</strong>s inverse using (2.3.1):<br />

(4.1.6) p(ξ (n)<br />

i ) =<br />

and obtain<br />

(4.1.7) K −1<br />

n−1 <br />

j=0<br />

cj uj(ξ (n)<br />

i ), 1 ≤ i ≤ n,<br />

n = {uj(ξ (n)<br />

i+1<br />

)} 0≤i≤n−1 .<br />

0≤j≤n−1<br />

.


Transforms 67<br />

It is evident that K −1<br />

n = D (1)<br />

n K t nD (2)<br />

n , where D (1) and D (2) are the diagonal matrices<br />

given respectively by diag{uj 2 w}0≤j≤n−1 and diag{1/w (n)<br />

i+1 }0≤i≤n−1. This shows<br />

that, after an appropriate normalization of the elements of the two bases, the inverse of<br />

Kn is given by <strong>it</strong>s transpose.<br />

Similar arguments can be applied to the analysis of transforms based on Gauss-<br />

Lobatto nodes. Here the space Pn is generated e<strong>it</strong>her by {uk}0≤k≤n or by { ˜ l (n)<br />

j }0≤j≤n.<br />

Hence, we introduce a discrete Fourier transform ˜ Kn : Pn → Pn, which corresponds<br />

to an (n + 1) × (n + 1) matrix { ˜ kij} 0≤i≤n . To get the first n − 1 coefficients we use<br />

0≤j≤n<br />

(2.3.7) and (3.5.1):<br />

(4.1.8) ci =<br />

1<br />

ui 2 w<br />

n<br />

j=0<br />

For the last coefficient we recall (3.8.12)<br />

(4.1.9) cn =<br />

p(η (n)<br />

j ) ui(η (n)<br />

j ) ˜w (n)<br />

j , 0 ≤ i ≤ n − 1.<br />

1<br />

un 2 w,n<br />

Conversely, thanks to (2.3.1), one has<br />

(4.1.10) p(η (n)<br />

i ) =<br />

n<br />

j=0<br />

n<br />

j=0<br />

p(η (n)<br />

j ) un(η (n)<br />

j ) ˜w (n)<br />

j .<br />

cj uj(η (n)<br />

i ), 0 ≤ i ≤ n.<br />

Finally, for the Laguerre Gauss-Radau case, we define ˜ Kn : Pn−1 → Pn−1 as described<br />

above. Then, we have by virtue of (2.3.7) and (3.6.1)<br />

(4.1.11) ci =<br />

1<br />

L (α)<br />

i 2 w<br />

(4.1.12) p(η (n)<br />

i ) =<br />

n−1 <br />

j=0<br />

n−1 <br />

j=0<br />

p(η (n)<br />

j ) L (α)<br />

i (η (n)<br />

j ) ˜w (n)<br />

j , 0 ≤ i ≤ n − 1,<br />

cj L (α)<br />

j (η (n)<br />

i ), 0 ≤ i ≤ n − 1.<br />

The scaled weights introduced in section 3.10 can be taken into account in the evaluation<br />

of (4.1.11).


68 Polynomial Approximation of Differential Equations<br />

Henceforth, depending on the basis used for the representation of a polynomial, we<br />

distinguish among the two cases w<strong>it</strong>h the help of the following terminology. When the<br />

space of polynomials is intended to be the span of the uk’s, then <strong>it</strong> is isomorph to the<br />

set of Fourier coefficients. In this s<strong>it</strong>uation we call <strong>it</strong> the frequency space. Otherwise,<br />

when the isomorphism is naturally established w<strong>it</strong>h the set of the point values, <strong>it</strong> will<br />

be denoted by physical space. There are s<strong>it</strong>uations in which <strong>it</strong> is more convenient to<br />

work w<strong>it</strong>h one of the two representations. The discrete Fourier transform allows passage<br />

from the physical to the frequency space, thus both representations can be used to their<br />

fullest advantage.<br />

The cost of implementing the transform or <strong>it</strong>s inverse is proportional to n 2 , since<br />

<strong>it</strong> corresponds to a matrix-vector multiplication. Most of this chapter is dedicated to<br />

the analysis of a special case which occurs when Chebyshev polynomials are used. For<br />

this choice of polynomials, fast algor<strong>it</strong>hms have been developed to perform the discrete<br />

Fourier transform, considerably reducing the computational expense.<br />

4.2 Aliasing<br />

Two operators Πw,n and Iw,n have been respectively introduced in sections 2.4 and<br />

3.3. For a continuous function f, both the images Πw,n−1f and Iw,nf are in Pn−1,<br />

which is interpreted as frequency or physical space respectively. Unless f is a polynomial<br />

in Pn−1, the projection and the interpolant do not coincide. We show two examples.<br />

In figures 4.2.1 and 4.2.2, we plot in [−1,1] the projection and the interpolant of the<br />

same degree, corresponding to two different functions. In the first figure we have set<br />

f(x) = 4|x| − 2 − 1, w(x) ≡ 1 (Legendre weight function) and n = 8. In the second<br />

we have set f(x) = 1<br />

2 − |sinπx|, w(x) = 1/ √ 1 − x 2 (Chebyshev weight function)<br />

and n = 13.


Transforms 69<br />

Figure 4.2.1 - Legendre projection and interpolant of degree 7 for the<br />

<br />

<br />

function f(x) = 4|x| − 2<br />

− 1, x ∈ [−1,1].<br />

Figure 4.2.2 - Chebyshev projection and interpolant of degree 12 for the<br />

function f(x) = 1<br />

2 − |sin πx|, x ∈ [−1,1].


70 Polynomial Approximation of Differential Equations<br />

Thus, <strong>it</strong> is natural to define, for n ≥ 1, the operator Aw,n = Iw,n − Πw,n−1. The<br />

polynomial in Pn−1<br />

(4.2.1) Aw,nf = Iw,nf − Πw,n−1f, n ≥ 1, f ∈ C 0 ([−1,1]),<br />

takes the name of aliasing error.<br />

Depending on the smoothness of f, this error tends to zero when n → +∞. The proof<br />

of this fact and the characterization of the aliasing error is given in section 6.6.<br />

Another type of aliasing error is obtained by setting<br />

(4.2.2) Ãw,nf = Ĩw,nf − Πw,nf, f ∈ C 0 ([−1,1]),<br />

where n ≥ 1. This is also studied later on.<br />

4.3 Fast Fourier transform<br />

Let us examine what happens by applying the discrete Fourier transform in the special<br />

case of the Chebyshev basis. By virtue of (1.5.6), (3.1.4), (3.4.6) and (2.2.12), the entries<br />

of Kn in (4.1.5) take the form<br />

(4.3.1) kij =<br />

(−1) i<br />

n<br />

<br />

1 if i = 0, 0 ≤ j ≤ n − 1<br />

i(2j + 1)π<br />

cos ×<br />

2n<br />

2 if 1 ≤ i ≤ n − 1, 0 ≤ j ≤ n − 1<br />

For the same reason, due to (3.1.11), (3.5.7) and (3.8.5), the entries of ˜ Kn are<br />

(4.3.2) ˜ kij =<br />

(−1) i<br />

n<br />

<br />

ijπ<br />

cos ×<br />

n<br />

⎧<br />

1<br />

2 if i = 0, j = 0 or j = n<br />

⎪⎨<br />

1<br />

1<br />

if i = 0, 1 ≤ j ≤ n − 1<br />

if 1 ≤ i ≤ n − 1, j = 0 or j = n<br />

2<br />

1 ⎪⎩ 2<br />

1<br />

if 1 ≤ i ≤ n − 1, 1 ≤ j ≤ n − 1<br />

if i = n, j = 0 or j = n<br />

if i = n, 1 ≤ j ≤ n − 1<br />

.<br />

.


Transforms 71<br />

The main ingredient in expressions (4.3.1) and (4.3.2) is a term of the form cos m1m2<br />

n π,<br />

where m1 and m2 are integers. Due to the periodic<strong>it</strong>y of the cosine function, the entries<br />

of Kn and ˜ Kn may attain the same values for different choices of i,j ∈ N. The<br />

most remarkable case is when n is of the form 2 k for some k ∈ N. In this s<strong>it</strong>uation<br />

a lot of coefficients are identical and regularly distributed in the matrix. The poorest<br />

case is when n is a prime number. The entries generated then attain a larger number<br />

of distinct values and substructures are not easily recognized. These observations are<br />

the starting point of an efficient algor<strong>it</strong>hm for the evaluation of the Fourier transform<br />

and <strong>it</strong>s inverse, w<strong>it</strong>h a computational cost less than that needed for a matrix-vector<br />

multiplication. This procedure is known as Fast Fourier Transform (abbreviated by<br />

FFT).<br />

The original version of the FFT was officially introduced in cooley and tukey(1965),<br />

although the concept had already been published. Fast Fourier transforms are best<br />

defined in complex space. On the other hand, the transform involving the cosine can be<br />

interpreted as a type of projection into the set of the real numbers. A lot of variants and<br />

improvements have been added by various authors. At the same time, computer pro-<br />

grams have been coded, w<strong>it</strong>h the aim of giving a fast and reliable product. High qual<strong>it</strong>y<br />

versions of the FFT based on modern parallel arch<strong>it</strong>ectures are also available. Applica-<br />

tions in the framework of spectral computations for parallel processors are examined in<br />

pelz(1990).<br />

The FFT algor<strong>it</strong>hm is a l<strong>it</strong>tle b<strong>it</strong> tricky and needs a certain amount of perseverance<br />

in order to be fully understood. Many books explain, using numerous examples, the<br />

basic ideas of the FFT as well as series of generalizations. Usually, they start w<strong>it</strong>h<br />

the complex formulation restricted to the case n = 2 k . Other generalizations (known<br />

as multi-radix FFT algor<strong>it</strong>hms) apply when n is not a power of 2, giving however less<br />

efficient performances. Among the texts we suggest for instance brigham(1974), bini,<br />

capovani, lotti and romani(1981), nussbaumer(1981). Since a full discussion of the<br />

FFT implementation is beyond the scope of this book, we only present an elementary<br />

expos<strong>it</strong>ion to bring out <strong>it</strong>s essential features.<br />

The first problem we have in mind is to determine a fast way to compute matrix-


72 Polynomial Approximation of Differential Equations<br />

vector multiplications w<strong>it</strong>h (4.3.1). As we mentioned before, <strong>it</strong> is convenient to operate<br />

w<strong>it</strong>h complex numbers. Thus, let i denote the complex un<strong>it</strong>y, i.e., i 2 = −1. Then, we<br />

consider the transform<br />

(4.3.3) γm =<br />

n−1 <br />

j=0<br />

δj e 2imjπ/n , 0 ≤ m ≤ n − 1.<br />

Here γm, 0 ≤ m ≤ n − 1, and δj, 0 ≤ j ≤ n − 1, are complex variables, related by<br />

the standard complex FFT. We study later how to efficiently compute (4.3.3). We now<br />

establish a relation between (4.3.3) and (4.1.4) in the Chebyshev case. To this end,<br />

the papers of cooley, lewis and welsh(1970), and brachet et al.(1983) give useful<br />

indications. Assume that n is even. We first define the input data as follows:<br />

(4.3.4) δj<br />

⎧<br />

⎨p(ξ<br />

:=<br />

⎩<br />

(n)<br />

2j+1 ) n 0 ≤ j ≤ 2 − 1,<br />

≤ j ≤ n − 1.<br />

p(ξ (n)<br />

2n−2j )<br />

W<strong>it</strong>h this choice, we compute γm, 0 ≤ m ≤ n − 1. Finally, one verifies that<br />

(4.3.5) cm = (−1)m<br />

n<br />

·<br />

γ0<br />

n<br />

2<br />

if m = 0,<br />

γm e imπ/2n + γn−m e −imπ/2n if 1 ≤ m ≤ n − 1.<br />

Of course, this final computation has a cost only proportional to n. The procedure can<br />

be further improved by virtue of the analysis presented in swarztrauber(1986). Since<br />

the δj’s are real, one can spl<strong>it</strong> the data in a su<strong>it</strong>able way, to allow the computation of<br />

two complex FFTs of length n<br />

2<br />

instead of n. We skip the details for simplic<strong>it</strong>y.<br />

The inverse transform (see (4.1.6)) is clearly treated in a very similar way.<br />

The next problem is related to matrix-vector multiplications involving (4.3.2). We<br />

now define<br />

(4.3.6) δj :=<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

p(η (n)<br />

0 ) j=0,<br />

p(η (n)<br />

2j ) + i<br />

<br />

p(η (n)<br />

2j+1 ) − p(η(n) 2j−1 )<br />

<br />

1 ≤ j ≤ n<br />

2<br />

p(η (n)<br />

n ) j = n<br />

2 ,<br />

p(η (n)<br />

2n−2j ) + i<br />

<br />

p(η (n)<br />

2n−2j−1 ) − p(η(n) 2n−2j+1 )<br />

<br />

n<br />

2<br />

− 1,<br />

+ 1 ≤ j ≤ n − 1.


Transforms 73<br />

After evaluating the γm’s by (4.3.3), the output data can be recovered by<br />

(4.3.7)<br />

cm = (−1)m<br />

2n ·<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

p(η (n)<br />

0 ) + p(η (n)<br />

n ) + 2 n−1<br />

j=1 p(η(n)<br />

j ) m = 0,<br />

γm<br />

<br />

1 + 2sin πm<br />

n<br />

<br />

−1<br />

+ γn−m 1 − 2sin πm<br />

<br />

−1<br />

n<br />

p(η (n)<br />

0 ) + p(η (n)<br />

n ) + 2 n−1<br />

j=1 (−1)j p(η (n)<br />

j ) m = n.<br />

1 ≤ m ≤ n − 1,<br />

Here the upper bar denotes complex conjugate. The cumbersome verification of (4.3.7)<br />

is left to the reader. We note that the output coefficients in (4.3.7) are real. Other sug-<br />

gestions and improvements are contained for instance in swarztrauber(1986). Again,<br />

once the γm’s have been determined, the cost to implement (4.3.7) is proportional to<br />

n.<br />

We proceed now w<strong>it</strong>h the analysis of relation (4.3.3) and we consider the special<br />

case when n = 8 = 2 3 . The matrix corresponding to the linear transformation takes the<br />

form<br />

(4.3.8) Φ =<br />

⎡<br />

⎢<br />

⎣<br />

φ0 φ0 φ0 φ0 φ0 φ0 φ0 φ0 φ0 φ1 φ2 φ3 φ4 φ5 φ6 φ7 φ0 φ2 φ4 φ6 φ0 φ2 φ4 φ6 φ0 φ3 φ6 φ1 φ4 φ7 φ2 φ5 φ0 φ4 φ0 φ4 φ0 φ4 φ0 φ4 φ0 φ5 φ2 φ7 φ4 φ1 φ6 φ3 φ0 φ6 φ4 φ2 φ0 φ6 φ4 φ2 φ0 φ7 φ6 φ5 φ4 φ3 φ2 φ1 where φ := e iπ/4 and we noted that φ k+8 = φ k , ∀k ∈ N.<br />

Moreover, the linear system can be also wr<strong>it</strong>ten as<br />

⎡ ⎤<br />

⎡<br />

(4.3.9)<br />

⎢<br />

⎣<br />

γ0<br />

γ4<br />

γ2<br />

γ6<br />

γ1<br />

γ5<br />

γ3<br />

γ7<br />

⎥<br />

⎦<br />

= Φ1 Φ2 Φ3<br />

where the matrices Φk, 1 ≤ k ≤ 3, are given by<br />

⎢<br />

⎣<br />

δ0<br />

δ1<br />

δ2<br />

δ3<br />

δ4<br />

δ5<br />

δ6<br />

δ7<br />

⎤<br />

⎥<br />

⎥,<br />

⎥<br />

⎦<br />

⎤<br />

⎥<br />

⎥,<br />

⎥<br />


74 Polynomial Approximation of Differential Equations<br />

(4.3.10) Φ1 =<br />

(4.3.11) Φ2 =<br />

(4.3.12) Φ3 =<br />

⎡<br />

⎢<br />

⎣<br />

⎡<br />

⎢<br />

⎣<br />

⎡<br />

⎢<br />

⎣<br />

φ0 φ0 0 0 0 0 0 0<br />

φ0 φ4 0 0 0 0 0 0<br />

0 0 φ0 φ2 0 0 0 0<br />

0 0 φ0 φ6 0 0 0 0<br />

0 0 0 0 φ0 φ1 0 0<br />

0 0 0 0 φ0 φ5 0 0<br />

0 0 0 0 0 0 φ0 φ3 0 0 0 0 0 0 φ0 φ7 φ0 0 φ0 0 0 0 0 0<br />

0 φ0 0 φ0 0 0 0 0<br />

φ0 0 φ4 0 0 0 0 0<br />

0 φ0 0 φ4 0 0 0 0<br />

0 0 0 0 φ0 0 φ2 0<br />

0 0 0 0 0 φ0 0 φ2 0 0 0 0 φ0 0 φ6 0<br />

0 0 0 0 0 φ0 0 φ6 φ0 0 0 0 φ0 0 0 0<br />

0 φ0 0 0 0 φ0 0 0<br />

0 0 φ0 0 0 0 φ0 0<br />

0 0 0 φ0 0 0 0 φ0 φ0 0 0 0 φ4 0 0 0<br />

0 φ0 0 0 0 φ4 0 0<br />

0 0 φ0 0 0 0 φ4 0<br />

0 0 0 φ0 0 0 0 φ4 This spl<strong>it</strong>ting is the kernel of the FFT algor<strong>it</strong>hm. In general, starting from the n×n full<br />

matrix Φ (where n is a power of 2), we can find a su<strong>it</strong>able sequence of n × n matrices<br />

Φk, 1 ≤ k ≤ log 2n, whose product Φ ∗ := log 2 n<br />

k=1 Φk allows the recovery of Φ. As the<br />

reader can observe, the data entries in the left-hand side of (4.3.9) have been permuted.<br />

Therefore, Φ is obtained after a su<strong>it</strong>able recombination of the rows of Φ ∗ .<br />

The new matrices are very sparse. Actually, a vector multiplication just involves<br />

two nonzero entries for each row. The FFT algor<strong>it</strong>hm takes advantage of the reg-<br />

ular structure of the Φk’s. This reduces the matrix-vector multiplication for any<br />

⎤<br />

⎥<br />

⎥,<br />

⎥<br />

⎦<br />

⎤<br />

⎥<br />

⎥,<br />

⎥<br />

⎦<br />

⎤<br />

⎥<br />

⎥.<br />

⎥<br />


Transforms 75<br />

Φk, 1 ≤ k ≤ log 2n, to a procedure whose cost is only proportional to n. Consequently,<br />

the complete FFT can be performed w<strong>it</strong>h a cost proportional to nlog 2n, provided n is<br />

a power of 2. This shows that, for large values of n, the procedure is efficient. Whether<br />

this approach is compet<strong>it</strong>ive for small values of n, depends strongly on code implemen-<br />

tation. Performance is also affected by different computer arch<strong>it</strong>ectures.<br />

A clear description, both intu<strong>it</strong>ive and rigorous, of the steps to be followed to minimize<br />

the operations, is given in brigham(1974), where a flow chart is presented. General-<br />

izations to the case where n is not a power of 2, are discussed e.g. in singleton(1969)<br />

and temperton(1983).<br />

Finally, we touch on the problem of reordering the output data, known as un-<br />

scrambling procedure. Consider for example equation (4.3.9). We associate to any index<br />

0 ≤ m ≤ 7, <strong>it</strong>s 3-dig<strong>it</strong> binary representation, i.e.<br />

0 000 0<br />

1 001 4<br />

2 010 2<br />

3 011 6<br />

4 100 1<br />

5 101 5<br />

6 110 3<br />

7 111 7<br />

The last column shows the indices obtained by reading the binary numbers from right to<br />

left. These exactly match the indices of the left-hand side vector in (4.3.9). Reordering<br />

this last set results in the reordering of the output data. The rough outline above can<br />

be easily generalized. Recent variants of the FFT algor<strong>it</strong>hm (self-sorting FFT) do not<br />

require unscrambling of the data after applying the transform. We refer the reader to<br />

specialized publications for further information.


76 Polynomial Approximation of Differential Equations<br />

4.4 Other fast methods<br />

There are s<strong>it</strong>uations in which a polynomial, defined by the values assumed at the nodes<br />

of some Gauss integration formula, must be determined at the points corresponding to<br />

the relative Gauss-Lobatto formula. More precisely, from (3.2.2), we get for p ∈ Pn−1<br />

n<br />

(4.4.1) p(η (n)<br />

i ) =<br />

j=1<br />

Conversely, if p ∈ Pn, we obtain by (3.2.7)<br />

n<br />

(4.4.2) p(ξ (n)<br />

i ) =<br />

j=0<br />

p(ξ (n)<br />

j ) l (n)<br />

j (η (n)<br />

i ), 0 ≤ i ≤ n.<br />

p(η (n)<br />

j ) ˜ l (n)<br />

j (ξ (n)<br />

i ), 1 ≤ i ≤ n.<br />

As usual, particularly interesting is the Chebyshev case, where this interpolation process<br />

can be optimized. By (4.1.4) or (4.1.8)-(4.1.9), we recover the Fourier coefficients using<br />

the FFT. Then, in place of (4.4.1), one computes<br />

(4.4.3) p(η (n)<br />

i ) =<br />

n−1 <br />

j=0<br />

cj Tj(η (n)<br />

n−1 <br />

i ) = −<br />

Furthermore, in place of (4.4.2), we can wr<strong>it</strong>e<br />

n<br />

(4.4.4) p(ξ (n)<br />

i ) =<br />

j=0<br />

cj Tj(ξ (n)<br />

i ) = −<br />

so that an FFT-like algor<strong>it</strong>hm can still be used.<br />

n<br />

j=0<br />

j=0<br />

cj cos<br />

cj cos ij<br />

π, 0 ≤ i ≤ n.<br />

n<br />

j(2i − 1)<br />

π, 1 ≤ i ≤ n,<br />

2n<br />

The advantages of using two sets of points, like those considered above, in the dis-<br />

cretization of certain classes of partial differential equations, such as the incompressible<br />

Navier-Stokes equations (see section 13.4), have been investigated in malik, zang and<br />

hussaini(1985), and theoretically analyzed and extended in bernardi and maday<br />

(1988). The need for efficient transfer of data between the two grids, could favor the<br />

Chebyshev polynomials for these kind of computations.<br />

Add<strong>it</strong>ional algor<strong>it</strong>hms, based on the cosine FFT described in the previous section,<br />

have been adapted for other Jacobi polynomials (see orszag(1986)). Unfortunately,<br />

they are deduced from su<strong>it</strong>able asymptotic expansions. Therefore, they are fast and<br />

accurate only when n is qu<strong>it</strong>e large.


5<br />

FUNCTIONAL SPACES<br />

Before studying the convergence properties of orthogonal polynomials, <strong>it</strong> is necessary<br />

to specify the classes of functions to be approximated. This leads to the defin<strong>it</strong>ion of<br />

new functional spaces. An elementary introduction of the so called Sobolev spaces is<br />

given in this chapter. Beginners, w<strong>it</strong>h a weak mathematical background, can skip over<br />

this part in the first reading.<br />

5.1 The Lebesgue integral<br />

The Lebesgue integral extends the classical Riemann integral presented in the intro-<br />

ductory courses of calculus. This generalization allows the defin<strong>it</strong>ion of integrals of<br />

functions that belong to very large classes. This improvement does not only offer the<br />

possibil<strong>it</strong>y of dealing w<strong>it</strong>h pathological functions (most of which are not of interest to<br />

us), but provides new tools to help prove sophisticated results w<strong>it</strong>hin the framework of<br />

a simple and elegant theory.<br />

We start by reviewing the main properties of the Lebesgue measure theory. For a<br />

deeper analysis, we refer to the books by halmos(1971), kolmogorov and fomin(1961),<br />

hartman and mikusiński(1961), hildebrandt(1963), temple(1971), rao(1987), etc..


78 Polynomial Approximation of Differential Equations<br />

We only discuss the one dimensional case. Let I =]a,b[, a < b, be a bounded<br />

interval. The Lebesgue measure µ associated to a subset J ⊂<br />

Ī is a non negative<br />

real number. However, not all the subsets are measurable, though these cases are qu<strong>it</strong>e<br />

involved. Let us begin by considering very simple sets. Briefly, we review the most<br />

significant cases.<br />

The empty set J = ∅ has measure zero. If J ⊂<br />

Ī contains a single point or a<br />

fin<strong>it</strong>e number of points, then µ(J) := 0. When J is an interval w<strong>it</strong>h endpoints c and d,<br />

a ≤ c < d ≤ b, <strong>it</strong> is natural to require that<br />

µ(]c,d[) = µ([c,d[) = µ(]c,d]) = µ([c,d]) := d − c.<br />

When J is the union of a fin<strong>it</strong>e number m of disjoint subintervals Ki, 1 ≤ i ≤ m, then<br />

<strong>it</strong>s measure is given by µ(J) := m<br />

i=1 µ(Ki). Sets of this form are called elementary<br />

sets. Union, intersection and difference of two elementary sets is still an elementary set.<br />

The measure of an elementary set is easily reduced to the sum of lengths of one or more<br />

segments.<br />

A further generalization is to assume that J is the union of a countable collection<br />

of disjoint intervals Ki ⊂ Ī, i ∈ N; we then set µ(J) = ∞<br />

i=0 µ(Ki). This series<br />

converges since <strong>it</strong> is bounded by the value b−a. We note that, if the intervals Ki, i ∈ N,<br />

are not all disjoint (i.e., there exist i,j ∈ N such that Ki ∩Kj = ∅), then we can replace<br />

the family {Ki}i∈N by the family {K ∗ i }i∈N such that J = ∪ ∞ i=0 K∗ i and the K∗ i ’s<br />

are disjoint. In this case, we have the inequal<strong>it</strong>y<br />

(5.1.1) µ(J) =<br />

∞<br />

i=0<br />

µ(K ∗ i ) ≤<br />

∞<br />

i=0<br />

µ(Ki).<br />

Now, let J ⊂ Ī be a generic subset. Inspired by (5.1.1), we can always define the outer<br />

measure µ ∗ of J. This is given by<br />

(5.1.2) µ ∗ <br />

∞<br />

<br />

∞<br />

(J) := inf µ(Ki) <br />

J ⊂<br />

i=0<br />

where the lower bound is taken over all the coverings of J by countable families of<br />

intervals Ki, i ∈ N. For elementary sets the outer measure µ ∗ coincides w<strong>it</strong>h the actual<br />

measure µ.<br />

i=0<br />

Ki<br />

<br />

,


Functional Spaces 79<br />

We are now ready to give the defin<strong>it</strong>ion of a Lebesgue measurable set. The subset J ⊂ Ī<br />

is measurable if and only if<br />

(5.1.3) µ ∗ (J) = µ( Ī) − µ∗ ( Ī − J).<br />

Then we set µ(J) := µ ∗ (J). The idea is the following. The term on the left-hand<br />

side of (5.1.3) is obtained by approximating J from outside w<strong>it</strong>h families of intervals.<br />

The term on the right-hand side represents in some sense an approximation from inside.<br />

When both the estimates coincide, then J is measurable.<br />

It is possible to show that the union and intersection of a fin<strong>it</strong>e number of measurable<br />

sets are still measurable. Moreover, and this is the main improvement w<strong>it</strong>h respect to<br />

Riemann measure theory, union and intersection of a countable family of measurable<br />

sets are still measurable. If the sets Ji, i ∈ N, are measurable and disjoint, then<br />

(5.1.4) µ<br />

∞<br />

i=0<br />

Ji<br />

<br />

=<br />

∞<br />

i=0<br />

µ(Ji).<br />

For instance, the set of rational numbers J = Q ∩Ī, is not Riemann measurable. Being<br />

the union of a countable collection of points, J is Lebesgue measurable and by (5.1.4)<br />

µ(J) = 0. If I is an unbounded interval, then J ⊂<br />

measurable for any bounded interval K ⊂<br />

also take the value +∞.<br />

Ī is measurable when J ∩ K is<br />

Ī. In this s<strong>it</strong>uation we assume that µ can<br />

We can now introduce the concept of measurable function. A function f : Ī → R<br />

is measurable when the set {x ∈<br />

Ī| f(x) < γ} is measurable for any γ ∈ R. It is<br />

possible to prove that the sum and the product of two measurable functions are still<br />

measurable.<br />

A measurable function that assumes a fin<strong>it</strong>e or countable set of values in<br />

Ī, is called<br />

simple function. Any bounded measurable function f can be represented as a uniform<br />

lim<strong>it</strong> of simple functions fn, n ∈ N (this means that limn→∞ sup x∈I |fn(x)−f(x)| = 0).


80 Polynomial Approximation of Differential Equations<br />

Two measurable functions f and g coincide almost everywhere (abbreviated by a.e.),<br />

when the measure of the set {x ∈ Ī| f(x) = g(x)} is zero. In this s<strong>it</strong>uation f and g<br />

are called equivalent.<br />

We have now collected enough ingredients to define the Lebesgue integral of a<br />

measurable function. Let us start w<strong>it</strong>h the integral of a simple function f. In this case,<br />

f attains at most a countable set of values γi, i ∈ N, and the sets Ai := {x ∈ Ī| f(x) =<br />

γi}, i ∈ N, are measurable. Then we have<br />

(5.1.5)<br />

<br />

I<br />

f dx :=<br />

provided the series is absolutely convergent.<br />

∞<br />

i=0<br />

γi µ(Ai),<br />

In general, let f be a measurable function; then there exists a sequence of simple func-<br />

tions fn, n ∈ N, uniformly convergent to f. Thus we set<br />

(5.1.6)<br />

<br />

I<br />

f dx := lim<br />

n→∞<br />

<br />

fn dx,<br />

I<br />

provided the lim<strong>it</strong> exists. In this case, the lim<strong>it</strong> does not depend on the approximating<br />

sequence {fn}n∈N.<br />

The sum of two integrable functions is an integrable function. Moreover, for continuous<br />

or monotone functions, the Lebesgue and Riemann integrals coincide.<br />

We have the implication<br />

(5.1.7)<br />

<br />

I<br />

|f| dx = 0 ⇐⇒ f ≡ 0 a.e.<br />

This shows that point values, achieved on sets whose measure is zero, are not meaningful<br />

for the evaluation of the integral. This observation is crucial. In the following, besides<br />

a given integrable function f, we shall not distinguish among functions equivalent to<br />

f. This is because we shall work w<strong>it</strong>h norms of f based on the Lebesgue integral, and<br />

these are not affected by modifications of f on a set of measure zero.


Functional Spaces 81<br />

5.2 Spaces of measurable functions<br />

In section 2.1, we introduced an inner product and a norm in the space of continuous<br />

functions (see (2.1.8) and (2.1.9)). As we shall see in the sequel, C0 ( Ī) is not a su<strong>it</strong>able<br />

space for this kind of norm. Besides, in view of the chosen applications, we must to<br />

enrich our space w<strong>it</strong>h new functions. Using the approach of section 2.1, this is qu<strong>it</strong>e<br />

cumbersome. For instance, including discontinuous functions is not straightforward.<br />

One of the reasons is that property (2.1.3) could not be satisfied in an extended space<br />

(for example, take u vanishing in I, w<strong>it</strong>h the exception of one point).<br />

Here, we would like to make this extension in the very large family of measurable<br />

functions. To this purpose, for any measurable function u, we shall denote by Cu the<br />

class of all the functions equivalent to u. Let w : I → R, be a pos<strong>it</strong>ive continuous<br />

weight function (we prefer to avoid generalizations concerning w). Then, we define the<br />

following functional space:<br />

(5.2.1) L 2 w(I) :=<br />

<br />

Cu<br />

<br />

<br />

u is measurable and<br />

<br />

I<br />

u 2 <br />

wdx < +∞ .<br />

Although the elements of L 2 w(I) are not functions, we shall not distinguish between u<br />

and Cu, since the integral in (5.2.1) does not recognize the different representatives of a<br />

certain class. So that, due to (5.1.7), we get the implication<br />

(5.2.2) u ∈ L 2 <br />

w(I) and u 2 wdx = 0 ⇐⇒ u ≡ 0.<br />

I<br />

The correct interpretation of the right-hand side in (5.2.2) is that a generic representative<br />

of the class of functions associated w<strong>it</strong>h u is an element of the class of functions vanishing<br />

almost everywhere. When I is bounded, the inclusion C 0 ( Ī) ⊂ L2 w(I) is similarly<br />

justified.<br />

In L 2 w(I) we can define an inner product and a norm which are the natural extensions<br />

of those given in section 2.1, i.e.<br />

(5.2.3) (u,v) L 2 w (I) :=<br />

<br />

I<br />

uv wdx, ∀u,v ∈ L 2 w(I),


82 Polynomial Approximation of Differential Equations<br />

(5.2.4) u L 2 w (I) :=<br />

<br />

I<br />

u 2 1<br />

wdx<br />

2<br />

, ∀u ∈ L 2 w(I).<br />

The property (2.1.3) is now exactly expressed by (5.2.2).<br />

Another functional space is related to the norm introduced in section 2.5. We first<br />

give some defin<strong>it</strong>ions. We say that a measurable function u is essentially bounded when<br />

we can find a constant γ ≥ 0 such that |u| ≤ γ almost everywhere. An essentially<br />

bounded function u behaves like a bounded function, except on sets of measure zero.<br />

Thus, we set<br />

(5.2.5) L ∞ (I) :=<br />

The corresponding norm is given by<br />

<br />

Cu<br />

<br />

<br />

<br />

u is essentially bounded .<br />

<br />

<br />

(5.2.6) uL∞ (I) := inf γ ≥ 0<br />

|u| ≤ γ a.e. .<br />

When I is bounded, continuous functions in Ī belong to L∞ (I). In this case, the norm<br />

(5.2.6) is equal to the norm · ∞, presented in section 2.5.<br />

Introductory discussions on the spaces analyzed herein, and their generalizations, are<br />

contained e.g. in vulikh(1963), goffman and pedrick(1965), okikiolu (1971), and<br />

wouk(1979). Other references are given in the following sections.<br />

5.3 Completeness<br />

We point out an important property of the functional spaces introduced above. We first<br />

recall that a sequence {un}n∈N in a normed space X is a Cauchy sequence when, for<br />

any ǫ > 0, we can find N ∈ N such that<br />

(5.3.1) un1 − un2 < ǫ, ∀n1,n2 > N.


Functional Spaces 83<br />

Every convergent sequence in X is a Cauchy sequence. The converse depends on the<br />

structure of X. This justifies the following defin<strong>it</strong>ion. A space X, where every Cauchy<br />

sequence converges to an element of X, is called complete space.<br />

Let I be a bounded interval. The space X = C 0 ( Ī) w<strong>it</strong>h the norm · w given in<br />

(2.1.9), is not complete. For instance, let Ī = [0,1] and un(x) = xn , x ∈ Ī, n ∈ N.<br />

Therefore, we obtain a Cauchy sequence in C 0 ( Ī) converging in the norm · w, w ≡ 1,<br />

to a discontinuous function.<br />

The correct norm to be used in C0 ( Ī) is defined by (2.5.1). This leads to a complete<br />

space, i.e., Cauchy sequences in C 0 ( Ī) w<strong>it</strong>h the norm · ∞ always converge to a<br />

continuous function in that norm (the well-known uniform convergence). Moreover,<br />

L ∞ (I) is also complete w<strong>it</strong>h the norm given in (5.2.6). In the following we set (see<br />

(5.2.6))<br />

(5.3.2) u C 0 (Ī) := u L ∞ (I), ∀u ∈ C 0 ( Ī),<br />

provided e<strong>it</strong>her I or u are bounded.<br />

Another interesting result is that L 2 w(I) is complete w<strong>it</strong>h the norm given in (5.2.4).<br />

When I is bounded, L2 w(I) is the completion of C0 ( Ī) w<strong>it</strong>h the norm (2.1.9). This means<br />

that any function in L 2 w(I) can be approximated in the norm of L 2 w(I) by a sequence<br />

of continuous functions. In this case, not only does C 0 ( Ī) ⊂ L2 w(I), but there exists a<br />

constant K > 0 (depending on µ(I) and w) such that<br />

(5.3.3) u L 2 w (I) ≤ K u C 0 (Ī), ∀u ∈ C 0 ( Ī).<br />

A complete space, whose norm is related to an inner product, is called a Hilbert space.<br />

In particular L2 w(I) is a Hilbert space, C0 ( Ī) is not a Hilbert space. More results are<br />

presented in the classical texts of functional analysis.


84 Polynomial Approximation of Differential Equations<br />

5.4 Weak derivatives<br />

We recall that the space Ck ( Ī), where k ≥ 1 is an integer, is given by the functions<br />

whose derivatives, of order less or equal to k, exist and are continuous in Ī. Following<br />

section 5.3, when I is bounded, a norm for Ck ( Ī) is obtained by setting<br />

(5.4.1) u C k (Ī) := u C 0 (Ī) +<br />

k<br />

<br />

<br />

<br />

d<br />

<br />

mu dxm <br />

<br />

<br />

m=1<br />

C 0 (Ī)<br />

, u ∈ C k ( Ī), k ≥ 1.<br />

W<strong>it</strong>h this norm C k ( Ī), k ≥ 1, is a complete space. In particular, when u ∈ Ck ( Ī),<br />

∀k ∈ N, we say that u ∈ C∞ ( Ī). We are mainly concerned w<strong>it</strong>h a particular subspace<br />

of C∞ ( Ī). This is denoted by C∞ 0 (I) and consists of all the functions φ ∈ C∞ ( Ī) such<br />

that there exists a closed and bounded interval strictly included in I, outside of which φ<br />

vanishes. Therefore, since I is open, any function in C ∞ 0 (I) is zero in a neighborhood of<br />

the possible boundary points of I. Therefore, all the derivatives of a function in C ∞ 0 (I)<br />

still belong to C ∞ 0 (I).<br />

Up to this point, we only considered subspaces of the set of continuous functions.<br />

This is against the policy of expanding our functional spaces. We now give meaning<br />

to the differential calculus of measurable functions by introducing the concept of weak<br />

derivative. A Lebesgue integrable function u is differentiable in a weak sense, if there<br />

exists another integrable function v such that<br />

<br />

(5.4.2)<br />

uφ ′ <br />

dx = −<br />

I<br />

vφ dx, ∀φ ∈ C<br />

I<br />

∞ 0 (I).<br />

The function v is the weak derivative of u which is indicated using classical notation.<br />

This defin<strong>it</strong>ion is not in conflict w<strong>it</strong>h the usual one. Actually, we claim that, if u ∈ C 1 ( Ī),<br />

then v ∈ C0 ( Ī) is exactly the standard derivative of u. To prove this statement, we just<br />

integrate the left-hand side of (5.4.2) by parts. Recalling the vanishing cond<strong>it</strong>ions on φ,<br />

we obtain as requested<br />

<br />

(5.4.3)<br />

(u<br />

I<br />

′ − v)φ dx = 0, ∀φ ∈ C ∞ 0 (I) =⇒ v ≡ u ′ .<br />

The implication in (5.4.3) can be proven w<strong>it</strong>h a l<strong>it</strong>tle knowledge of basic calculus.


Functional Spaces 85<br />

Higher-order derivatives are similarly defined. The measurable function v is the k-th<br />

derivative of u if<br />

<br />

(5.4.4)<br />

I<br />

u dk <br />

φ<br />

dx = (−1)k vφ dx, ∀φ ∈ C<br />

dxk I<br />

∞ 0 (I), k ≥ 1.<br />

The weak derivative of a function u is not uniquely determined. However, two different<br />

derivatives are equivalent, hence they are in the same class Cv.<br />

In general, not all integrable functions adm<strong>it</strong> a weak derivative, unless we enlarge<br />

the space of possible derivatives. This extension leads to the defin<strong>it</strong>ion of the space of<br />

distributions, whose elements are often not even representable as functions (the typical<br />

example is the Dirac delta). For simplic<strong>it</strong>y, we skip this subject and refer the interested<br />

reader to the following authors: schwartz(1966), treves(1967), guelfand, graev<br />

and vilenkin(1970), and garnir, de wilde and schmets(1973).<br />

In the following sections, we study how to use weak derivatives to define add<strong>it</strong>ional<br />

functional spaces.<br />

5.5 Transformation of measurable functions in R<br />

Let C denote the set of complex numbers. A function u : R → C is measurable when<br />

both the real part Re(u) and imaginary part Im(u) are measurable. Then we define<br />

(5.5.1) L 2 <br />

(R;C) :=<br />

<br />

<br />

u measurable and<br />

<br />

[(Re(u)) 2 + (Im(u)) 2 <br />

]dx < +∞ .<br />

Cu<br />

The space in (5.5.1) is a Hilbert space w<strong>it</strong>h the (complex) inner product<br />

<br />

(5.5.2) (u,v) L2 (R;C) := u¯v dx, u,v ∈ L 2 (R;C),<br />

where ¯v denotes the complex conjugate of v.<br />

R<br />

Let us now consider the operator F : L 2 (R;C) → L 2 (R;C). This maps u ∈ L 2 (R;C)<br />

to the function defined by:<br />

R


86 Polynomial Approximation of Differential Equations<br />

(5.5.3) [Fu](t) := 1<br />

<br />

√ u(x) e<br />

2π R<br />

−<strong>it</strong>x dx ∈ L 2 (R;C),<br />

where i is the imaginary un<strong>it</strong>y and the equal<strong>it</strong>y holds in R almost everywhere. Since<br />

u ∈ L 2 (R;C) does not in general imply that u is integrable in R, a l<strong>it</strong>tle care is necessary<br />

when manipulating expression (5.5.3). W<strong>it</strong>h some technical reasoning, one verifies that<br />

the setting is correct after appropriate interpretation (see for instance rudin(1966)).<br />

The function Fu is called Fourier transform of u. The operator F su<strong>it</strong>ably generalizes<br />

the complex discrete Fourier transform (4.3.3).<br />

The following relation is well-known:<br />

(5.5.4) Fu L 2 (R;C) = u L 2 (R;C), ∀u ∈ L 2 (R;C).<br />

In other words, F is an isometry.<br />

Moreover, the inverse operator F −1 : L 2 (R;C) → L 2 (R;C), adm<strong>it</strong>s the representation<br />

(5.5.5) [F −1 u](t) = 1<br />

√ 2π<br />

<br />

u(x) e<br />

R<br />

<strong>it</strong>x dx, a.e. ∈ R, ∀u ∈ L 2 (R;C).<br />

Many other properties characterize the transformation F (see rudin(1966)). We just<br />

mention some of those useful in the following sections.<br />

If both the real and the imaginary part of u ∈ L 2 (R;C) have a weak derivative (see<br />

previous section) and u ′ ∈ L 2 (R;C), then the following relation can be established for<br />

almost every t ∈ R:<br />

(5.5.6) [Fu ′ ](t) = −<br />

i t<br />

√ 2π [Fu] (t).<br />

Therefore, setting v(t) := t[Fu](t), t ∈ R, we have<br />

(5.5.7) u ′ ∈ L 2 (R;C) ⇐⇒ v ∈ L 2 (R;C).<br />

Since t ∈ R diverges at infin<strong>it</strong>y at the same rate as ρ(t) := (1 + t 2 ) 1/2 , t ∈ R, (5.5.7)<br />

can be rewr<strong>it</strong>ten as<br />

(5.5.8) u ′ ∈ L 2 (R;C) ⇐⇒ ρFu ∈ L 2 (R;C).


Functional Spaces 87<br />

The use of ρ(t) in place of t will be understood later on. Now, due to (5.5.8) and a<br />

propos<strong>it</strong>ion proven in escobedo and kavian(1987), if w(x) = ex2, x ∈ R, one can<br />

deduce that<br />

(5.5.9) u ∈ L 2 w(R), u ′ ∈ L 2 w(R) ⇐⇒ ρF(u √ w) ∈ L 2 (R;C).<br />

More generally, for any integer k ≥ 1, when u is differentiable k times, we have<br />

(5.5.10)<br />

d m u<br />

dx m ∈ L2 w(R), 0 ≤ m ≤ k ⇐⇒ ρ k F(u √ w) ∈ L 2 (R;C).<br />

5.6 Sobolev spaces in R<br />

Let k ≥ 1 be an integer. We introduce the weighted Sobolev space of order k in R as<br />

follows (see adams(1975), bergh and löfström(1976), triebel(1978)):<br />

(5.6.1) H k <br />

w(R) :=<br />

<br />

<br />

u differentiable k times and<br />

Cu<br />

dmu dxm ∈ L2 <br />

w(R), for 0 ≤ m ≤ k .<br />

The derivative is intended in the weak sense and dm u<br />

dx m = u when m = 0. We also<br />

set H 0 w(R) := L 2 w(R). Recalling (5.2.3) and (5.2.4), an inner product and a norm are<br />

introduced in H k w(R), k ∈ N, namely<br />

(5.6.2) (u,v) H k w (R) :=<br />

(5.6.3) u H k w (R) :=<br />

k<br />

m=0<br />

<br />

k <br />

<br />

dmu dxm <br />

<br />

m=0<br />

m d u<br />

dxm , dmv dxm <br />

L2 w (R)<br />

, ∀u,v ∈ H k w(R),<br />

2<br />

L2 w (R)<br />

1<br />

2<br />

, ∀u ∈ H k w(R).<br />

W<strong>it</strong>h this norm, H k w(R), k ∈ N, is a Hilbert space. The higher is the index k and the<br />

smoother the functions belonging to H k w(R) will be.


88 Polynomial Approximation of Differential Equations<br />

From now on, let us assume that w(x) = ex2, x ∈ R, i.e. the inverse of the Herm<strong>it</strong>e<br />

weight function. A first characterization is immediately obtained by virtue of relation<br />

(5.5.10)<br />

(5.6.4) H k w(R) =<br />

<br />

Cu<br />

<br />

<br />

ρ k F(u √ w) ∈ L 2 <br />

(R;C) ,<br />

where ρ(t) = (1 + t 2 ) 1/2 , t ∈ R. This new point of view suggests the introduction of<br />

other Sobolev spaces. Namely, let s ≥ 0 be a real number, we define<br />

(5.6.5) H s w(R) :=<br />

<br />

Cu<br />

<br />

<br />

ρ s F(u √ w) ∈ L 2 <br />

(R;C) .<br />

We note that ρ s , s ≥ 0 is a smooth function in R. This justifies the use of ρ(t) in place<br />

of t.<br />

Of course, when s is integer, the two defin<strong>it</strong>ions (5.6.1) and (5.6.5) coincide. It is clear<br />

that we have the inclusion H s1<br />

w (R) ⊂ H s2<br />

w (R), if s1 > s2 ≥ 0. In this way, although we<br />

cannot speak of derivatives of order s when s is not an integer, we can instead decide<br />

when a function u belongs to H s w(R). In this case, u will be more regular than a function<br />

in H [s]<br />

w (R) and less regular than a function in H [s]+1<br />

w (R), where [s] denotes the integer<br />

part of s. For instance, let u(x) = |x|e−x2, x ∈ R. Then, one can easily check that<br />

u ∈ H 1 w(R). On the other hand, for any s such that 0 ≤ s < 1, we have u ∈ H s w(R).<br />

By (5.5.2), a norm in H s w(R), s ≥ 0, is defined by<br />

(5.6.6) |||u||| H s w (R) := ρ s F(u √ w) L 2 (R;C), ∀u ∈ H s w(R).<br />

One can prove that, when s is an integer, the norms (5.6.3) and (5.6.6) are equivalent (see<br />

triebel(1978), p.177). Actually, for any k ∈ N, we can find two constants C1,C2 > 0,<br />

such that<br />

(5.6.7) C1 u H k w (R) ≤ |||u||| H k w (R) ≤ C2u H k w (R), ∀u ∈ H k w(R).<br />

When s is not integer, <strong>it</strong> is possible to show (see triebel(1978), p.189) that the norm<br />

||| · ||| H s w (R) is equivalent to:


Functional Spaces 89<br />

(5.6.8) u H s w (R) := u H [s]<br />

w (R) +<br />

<br />

R×R<br />

|v(x) − v(y)| 2<br />

|x − y| 1+2σ 1<br />

2<br />

dxdy , ∀u ∈ H s w(R),<br />

where v := d[s]<br />

dx [s] (u √ w) and σ := s − [s].<br />

Depending on the context, we shall make use e<strong>it</strong>her of the representations (5.6.3) and<br />

(5.6.8) or of the expression (5.6.6). In all three cases, H s w(R) turns out to be a complete<br />

space.<br />

We can easily check that any function of the form pe−x2, where p is a polynomial,<br />

belongs to H s w(R), ∀s ≥ 0. At the same time, if u ∈ C k (R), k ∈ N, and u and <strong>it</strong>s<br />

derivatives of order less or equal to k decay faster than e −x2 /2 at infin<strong>it</strong>y, then we obtain<br />

u ∈ H k w(R). The converse of this statement leads to the Sobolev embedding theorem (see<br />

triebel(1978), p.206), which can be stated as follows. For any k ∈ N, if u ∈ H s w(R),<br />

w<strong>it</strong>h s > k + 1<br />

2 , then u ∈ Ck (R). Furthermore, a constant K > 0 exists such that<br />

(5.6.9) u √ w C k (R) ≤ K u H s w (R), ∀u ∈ H s w(R), s > k + 1<br />

2 .<br />

This result is significant since <strong>it</strong> relates abstract Sobolev spaces w<strong>it</strong>h the classical spaces<br />

of continuous and differentiable functions. To be precise, the inclusion H s w(R) ⊂ C 0 (R),<br />

for s > 1<br />

2 , means that any class Cu ∈ H s w(R) contains a continuous representative u.<br />

We also have (see funaro and kavian(1988))<br />

(5.6.10) ρ µ u √ wC0 (R) ≤ K uHs w (R), ∀u ∈ H s w(R), s > 1<br />

2 + µ and µ ≥ 0.<br />

An interesting property is that any function in H s w(R), s ≥ 0, can be approximated in<br />

the norm of H s w(R) by a sequence of functions in C ∞ 0 (R).<br />

Finally, we mention the following inequal<strong>it</strong>y. Let σ,τ ≥ 0 and u ∈ H σ w(R) ∩ H τ w(R),<br />

then we have u ∈ H s w(R), where s = (1−θ)σ+θτ, θ ∈ [0,1], and we can find a constant<br />

K > 0 such that<br />

(5.6.11) u H s w (R) ≤ K u 1−θ<br />

H σ w (R) uθ H τ w (R), ∀u ∈ Hσ w(R) ∩ H τ w(R).


90 Polynomial Approximation of Differential Equations<br />

5.7 Sobolev spaces in intervals<br />

For any integer k ≥ 1, the weighted Sobolev space of order k in I ⊂ R is defined by<br />

(5.7.1) H k <br />

w(I) :=<br />

<br />

<br />

u differentiable k times and<br />

Cu<br />

dmu dxm ∈ L2 <br />

w(I), for 0 ≤ m ≤ k .<br />

Here, w : I → R is a weight function. Moreover H 0 w(I) := L 2 w(I). For every<br />

k ∈ N, H k w(I) is a Hilbert space w<strong>it</strong>h the inner product and the norm<br />

(5.7.2) (u,v) H k w (I) :=<br />

(5.7.3) u H k w (I) :=<br />

k<br />

m=0<br />

m d u<br />

dxm , dmv dxm <br />

L2 w (I)<br />

, ∀u,v ∈ H k w(I),<br />

<br />

k <br />

<br />

dmu dxm <br />

<br />

m=0<br />

2<br />

L2 w (I)<br />

1<br />

2<br />

, ∀u ∈ H k w(I).<br />

Many fundamental properties about these spaces are given in kufner(1980). We recall<br />

that, for a function u ∈ H k w(I), there exists a sequence of regular functions {vn}n∈N<br />

(we can assume for example vn ∈ C∞ ( Ī), ∀n ∈ N), converging to u in the norm of<br />

H k w(I), i.e., limn→+∞ vn − u H k w (I) = 0. In the following, using a standard functional<br />

analysis technique, many theorems will be proven w<strong>it</strong>h the help of this property. The<br />

statement to be proven is checked for regular functions and successively extended to<br />

Sobolev spaces w<strong>it</strong>h a lim<strong>it</strong> procedure. For simplic<strong>it</strong>y, we shall often om<strong>it</strong> the details<br />

when applying this kind of technique.<br />

Let us examine the Jacobi weight function w(x) = (1−x) α (1+x) β , x ∈ I =]−1,1[,<br />

where the parameters satisfy the cond<strong>it</strong>ions −1 < α < 1 and −1 < β < 1. In this case,<br />

one verifies that H 1 w(I) ⊂ C 0 ( Ī). Moreover, we can find two constants K1, K2 > 0,<br />

such that<br />

(5.7.4) u L 2 w (I) ≤ K1 u C 0 (Ī) ≤ K2 u H 1 w (I), ∀u ∈ H 1 w(I).<br />

As a byproduct of Hardy’s inequal<strong>it</strong>y (see for instance lions and magenes(1972), p.49),<br />

one obtains the relation


Functional Spaces 91<br />

(5.7.5)<br />

<br />

u<br />

<br />

1 − x2 <br />

<br />

L 2 w (I)<br />

Now, we introduce another space, namely<br />

≤ K u ′ L 2 w (I), ∀u ∈ H 1 w(I) w<strong>it</strong>h u(±1) = 0.<br />

(5.7.6) H 1 <br />

0,w(I) := u ∈ H 1 <br />

<br />

w(I) u(±1) = 0 .<br />

The following norm will be adopted in H 1 0,w(I):<br />

(5.7.7) u H 1 0,w (I) := u ′ L 2 w (I), ∀u ∈ H 1 0,w(I).<br />

To show that this is a norm we have to check (2.1.3). Actually, by (5.7.5), the cond<strong>it</strong>ion<br />

u H 1 0,w (I) = 0 implies u ≡ 0. In practice, in H 1 0,w(I), <strong>it</strong> is sufficient to bound the<br />

derivative of u in order to control the function <strong>it</strong>self.<br />

As in the previous section, we would like to define Sobolev spaces H s w(I), where<br />

s ≥ 0 is a real number. There are several approaches. A first approach consists in<br />

constructing a su<strong>it</strong>able prolongation of the weight w and functions in L 2 w(I) to the whole<br />

domain R. Then we can provide a defin<strong>it</strong>ion involving the transformation F following<br />

the guideline of section 5.6. Another defin<strong>it</strong>ion can be based on a formula similar to<br />

(5.6.8). However, the common practice is to define H s w(I) as an intermediate space<br />

between H [s]<br />

w (I) and H [s]+1<br />

w (I), via a technique called interpolation (see for instance<br />

lions and magenes(1972)). For the sake of simplic<strong>it</strong>y, we shall not enter into the<br />

details of this procedure.<br />

The case w ≡ 1 has been extensively investigated by many authors. For a general<br />

weight function, we cannot expect the various defin<strong>it</strong>ions to match. The l<strong>it</strong>erature on<br />

this subject does not cover all the possible cases and the theory is often qu<strong>it</strong>e compli-<br />

cated. Results have recently been published in maday(1990) on the interpolation of<br />

Sobolev spaces w<strong>it</strong>h the Chebyshev weight function. Though the use of spaces obtained<br />

by interpolation seems to be a promising approach in view of the applications, w<strong>it</strong>h<br />

a few exceptions, one can easily end up w<strong>it</strong>h abstract intermediate spaces, w<strong>it</strong>hout a<br />

satisfactory characterization of their elements.


92 Polynomial Approximation of Differential Equations<br />

* * * * * * * * * * * * * *<br />

From the point of view of functional analysis, the brief presentation set forth in these<br />

pages is far from exhaustive. In our opinion, however, <strong>it</strong> contains sufficient material for<br />

an in depth understanding of the applications that follow. We refer the readers that are<br />

more concerned w<strong>it</strong>h theoretical issues, to the numerous publications in this field. Very<br />

attractive to applied mathematicians are, for instance, the text of brezis(1983) and<br />

the introductory essay of brezzi and gilardi(1987), where the functional spaces are<br />

rigorously treated, but w<strong>it</strong>hout the tiresome formalism of the more specialized books.


6<br />

RESULTS IN<br />

APPROXIMATION THEORY<br />

According to a theorem of Weierstrass, any continuous function in a bounded closed in-<br />

terval can be uniformly approximated by polynomials. From this celebrated statement,<br />

a large variety of sophisticated results have emerged. We review those that are most<br />

relevant to the analysis of spectral methods.<br />

6.1 The problem of best approximation<br />

We begin w<strong>it</strong>h a classical theorem in approximation theory.<br />

Theorem 6.1.1 (Weierstrass) - Let I be a bounded interval and let f ∈ C0 ( Ī). Then,<br />

for any ǫ > 0, we can find n ∈ N and pn ∈ Pn such that<br />

(6.1.1) |f(x) − pn(x)| < ǫ, ∀x ∈ Ī.<br />

Hints for the proof - We just give a brief sketch of the proof in the interval<br />

Ī = [0,1].<br />

We follow the guideline of the original approach presented in bernstein(1912b). Other<br />

techniques are discussed for instance in timan(1963), todd(1963) and cheney(1966).<br />

For sufficiently large n, the approximating polynomial is required to be of the form<br />

(6.1.2)<br />

n<br />

<br />

j n<br />

pn(x) := f x<br />

n j<br />

j (1 − x) n−j , x ∈ [0,1],<br />

j=0


94 Polynomial Approximation of Differential Equations<br />

which leads to<br />

(6.1.3) f(x) − pn(x) =<br />

n<br />

j=0<br />

<br />

j n<br />

f(x) − f x<br />

n j<br />

j (1 − x) n−j , x ∈ [0,1].<br />

The proof that this error tends to zero uniformly is now a technical exercise. One argues<br />

as follows. For any x ∈ [0,1], the sum in (6.1.2) is decomposed into two parts. In the<br />

first part, the indices j are such that |x − j/n| is su<strong>it</strong>ably small. Therefore, from<br />

the uniform continu<strong>it</strong>y of f, the term |f(x) − f(j/n)| will also be small. To handle<br />

the remaining part of the summation, we observe that the function n j n−j<br />

j x (1 − x)<br />

achieves <strong>it</strong>s maximum for x = j/n, and far away from this value <strong>it</strong> decays qu<strong>it</strong>e fast.<br />

By appropriately applying the above information, one can, after some manipulation,<br />

conclude the proof.<br />

At this point, the following question arises. Among all the polynomials of degree<br />

less or equal to a fixed integer n, find the one which best approximates, uniformly in<br />

Ī, a given continuous function f. In practice, w<strong>it</strong>h the notations of section 2.5, for any<br />

n ∈ N, we would like to study the existence of Ψ∞,n(f) ∈ Pn such that<br />

(6.1.4) f − Ψ∞,n(f)∞ = inf<br />

ψ∈Pn<br />

f − ψ∞.<br />

This problem adm<strong>it</strong>s a unique solution, though the proof is very involved. An extensive<br />

and general treatise on this subject is given for instance in timan(1963). The n-degree<br />

polynomial Ψ∞,n(f) is called the polynomial of best uniform approximation of f in Ī.<br />

As a consequence of theorem 6.1.1, one immediately obtains<br />

(6.1.5) lim<br />

n→+∞ f − Ψ∞,n(f)∞ = 0.<br />

One can try to characterize the polynomial of best approximation of a certain degree.<br />

Results in this direction are considered in timan(1963) and rivlin(1969) where Cheby-<br />

shev polynomials play a fundamental role. As an example, theorem 2.5.3 provides the<br />

polynomial that best approximates the zero function in [−1,1], in the subset of Pn given<br />

by the polynomials of the form x n + {lower degree terms}.


Results in Approximation Theory 95<br />

We now collect some result on the problem of best approximation. We begin w<strong>it</strong>h the<br />

celebrated theorem by Jackson (see for instance feinerman and newman(1974)).<br />

Theorem 6.1.2 (Jackson) - Let I =]−1,1[, then <strong>it</strong> is possible to find a constant C > 0<br />

such that, for any f ∈ C0 ( Ī), we have<br />

(6.1.6) f − Ψ∞,n(f)∞ ≤ C sup<br />

|x1−x2| 0 such that, for any f ∈ Ck ( Ī), we have<br />

k 1<br />

(6.1.7) f − Ψ∞,n(f)∞ ≤ C sup<br />

n<br />

|x1−x2| 0. Then, for any ǫ > 0, we can find n ∈ N and pn ∈ Pn such that<br />

(6.1.8) |f(x) − pn(x)| e −δx < ǫ, ∀x ∈ Ī.<br />

Theorem 6.1.5 - Let I = R and let f ∈ C 0 (R) satisfy limx→±∞ f(x)e −δx2<br />

for a given δ > 0. Then, for any ǫ > 0, we can find n ∈ N and pn ∈ Pn such that<br />

(6.1.9) |f(x) − pn(x)| e −δx2<br />

< ǫ, ∀x ∈ R.<br />

= 0,


96 Polynomial Approximation of Differential Equations<br />

6.2 Estimates for the projection operator<br />

As usual, we denote by {un}n∈N the sequence of orthogonal polynomials in Ī (where<br />

Ī = [−1,1],<br />

Ī = [0,+∞[ or Ī ≡ R), w<strong>it</strong>h respect to the weight function w (we refer in<br />

particular to those introduced in chapter one).<br />

Another best approximation problem can be formulated in terms of the norm · w<br />

introduced in (2.1.9). In this case, we wish to find Ψw,n(f) ∈ Pn such that<br />

(6.2.1) f − Ψw,n(f)w = inf<br />

ψ∈Pn<br />

where f is a given function in Ī.<br />

f − ψw,<br />

For the reader acquainted w<strong>it</strong>h the results of chapter five, the natural space to develop<br />

the theory is L 2 w(I). Under this assumption, we can still define the Fourier coefficients<br />

of f as<br />

(6.2.2) ck := (f,uk) L2 w (I)<br />

uk2 L2 w (I)<br />

, k ∈ N.<br />

We note that the integral <br />

I fukwdx is fin<strong>it</strong>e in virtue of the Schwarz inequal<strong>it</strong>y (2.1.7).<br />

In this formulation we can also include the cases when I is not bounded, corresponding<br />

respectively to Laguerre and Herm<strong>it</strong>e polynomials. Now, even if the functions in L 2 w(I)<br />

are not required to be bounded, the exponential decay of the weight w is sufficient<br />

to insure the existence of the coefficients in (6.2.2). The reader unfamiliar w<strong>it</strong>h these<br />

functional spaces, can draw conclusions by thinking in terms of Riemann integrals and<br />

continuous functions, and by replacing the norms · L 2 w (I) and · C 0 (Ī), by · w<br />

and · ∞ respectively.<br />

The projector Πw,n : L 2 w(I) → Pn, n ∈ N, is defined in the usual way (see section 2.4).<br />

The next propos<strong>it</strong>ion fully characterizes the solution of problem (6.2.1).<br />

Theorem 6.2.1 - For any f ∈ L 2 w(I) and any n ∈ N, there exists a unique polynomial<br />

Ψw,n(f) ∈ Pn that satisfies (6.2.1). Moreover Ψw,n(f) = Πw,nf.


Results in Approximation Theory 97<br />

Proof - A polynomial ψ ∈ Pn can be wr<strong>it</strong>ten in the form ψ = n<br />

k=0 dkuk, for some<br />

real coefficients dk, 0 ≤ k ≤ n. Minimizing f −ψ L 2 w (I), or equivalently f −ψ 2 L 2 w (I),<br />

requires the derivatives<br />

(6.2.3)<br />

+<br />

n<br />

k=0<br />

∂<br />

∂dj<br />

f − ψ 2 L2 <br />

∂<br />

w (I) = f<br />

∂dj<br />

2 L2 w<br />

d 2 kuk 2 L 2 w (I)<br />

<br />

(I) − 2<br />

n<br />

k=0<br />

dk(f,uk) L 2 w (I)<br />

= −2 (f,uj) L 2 w (I) + 2dj uj 2 L 2 w (I),<br />

0 ≤ j ≤ n.<br />

We deduce that the unique minimum is attained when dj = cj, 0 ≤ j ≤ n, where the<br />

cj’s are the Fourier coefficients of f (see (6.2.2)). This ends the proof.<br />

In short, we can wr<strong>it</strong>e<br />

(6.2.4) f − Πw,nf L 2 w (I) = inf<br />

ψ∈Pn<br />

f − ψ L 2 w (I).<br />

Another interesting characterization is given in the following theorem.<br />

Theorem 6.2.2 - For any f ∈ L 2 w(I) and n ∈ N, we have<br />

(6.2.5)<br />

<br />

I<br />

(f − Πw,nf)φw dx = 0, ∀φ ∈ Pn.<br />

Proof - We fix φ ∈ Pn and define G : R → R by<br />

G(ν) := f − Πw,nf + νφ 2 L 2 w (I).<br />

We know from theorem 6.2.1 that ν = 0 is a minimum for G. Therefore<br />

(6.2.6) G ′ (ν) = 2<br />

<br />

(f − Πw,nf)φw dx + 2νφ<br />

I<br />

2 L2 w (I), ∀ν ∈ R.<br />

Imposing that G ′ (0) = 0, we get (6.2.5).


98 Polynomial Approximation of Differential Equations<br />

By virtue of this theorem, the operator Πw,n takes the name of orthogonal projector,<br />

since the error f − Πw,nf is orthogonal to the space Pn. Choosing in particular<br />

φ = Πw,nf in (6.2.5), application of the Schwarz inequal<strong>it</strong>y leads to<br />

(6.2.7) Πw,nf L 2 w (I) ≤ f L 2 w (I), ∀n ∈ N, ∀f ∈ L 2 w(I).<br />

The counterpart of relation (6.1.5) is deduced from the next result.<br />

Theorem 6.2.3 - Let I be bounded, then for any f ∈ L 2 w(I) , we have<br />

(6.2.8) lim<br />

n→+∞ f − Πw,nf L 2 w (I) = 0.<br />

Proof - If f ∈ C0 ( Ī), (6.2.4) and (5.3.3) together imply that<br />

(6.2.9) f −Πw,nf L 2 w (I) ≤ f −Ψ∞,n(f) L 2 w (I) ≤ f −Ψ∞,n(f) C 0 (Ī)<br />

The last term in (6.2.9) tends to zero in view of (6.1.5).<br />

<br />

I<br />

1<br />

2<br />

w dx .<br />

Now, we prove the statement for a general f ∈ L 2 w(I). We can find a sequence of<br />

functions {fm}m∈N, w<strong>it</strong>h fm ∈ C 0 ( Ī), m ∈ N, converging to f in the L2 w(I) norm.<br />

Therefore, using the triangle inequal<strong>it</strong>y, we first note that<br />

(6.2.10) f − Πw,nf L 2 w (I) ≤ f − fm L 2 w (I)<br />

+fm − Πw,nfm L 2 w (I) + Πw,n(fm − f) L 2 w (I), ∀m ∈ N.<br />

Finally, each one of the three terms on the right-hand side of (6.2.10) tends to zero<br />

when m and n grow (for the last term we can recall (6.2.7)).<br />

When I is not bounded, the proof of the previous theorem does not apply anymore.<br />

Nevertheless, the expression (6.2.8) holds also in the case of Laguerre and Herm<strong>it</strong>e<br />

polynomials. The proof of this fact is more delicate and we refer to courant and<br />

hilbert(1953), Vol.1, page 95, for the details. This means that, for all the orthogonal<br />

systems of polynomials here considered, one is allowed to wr<strong>it</strong>e


Results in Approximation Theory 99<br />

(6.2.11) f =<br />

∞<br />

ckuk, ∀f ∈ L 2 w(I),<br />

k=0<br />

where the ck’s are the Fourier coefficients of f and the equal<strong>it</strong>y has to be intended<br />

almost everywhere. By orthogonal<strong>it</strong>y, one also obtains the Parseval ident<strong>it</strong>y<br />

∞<br />

∀f ∈ L2w(I). (6.2.12) f 2 L2 w (I) =<br />

k=0<br />

c 2 k uk 2 L 2 w (I),<br />

From (6.2.11) and (6.2.12), one easily deduces that<br />

(6.2.13) f − Πw,nf L 2 w (I) ≤ f L 2 w (I), ∀f ∈ L 2 w(I).<br />

We remark that the convergence of the series (6.2.11) is not in general uniform in<br />

I. On the other hand, discontinuous functions can be also approximated. For instance,<br />

when f is continuous w<strong>it</strong>h the exception of the point x0 ∈] − 1,1[, then for Legendre<br />

expansions one has<br />

(6.2.14)<br />

∞<br />

k=0<br />

provided the two lim<strong>it</strong>s exist.<br />

ckuk(x0) = 1<br />

<br />

2<br />

lim<br />

x→x −<br />

0<br />

f(x) + lim<br />

x→x +<br />

f(x)<br />

0<br />

Figure 6.2.1 - Legendre orthogonal Figure 6.2.2 - Legendre orthogonal<br />

projections for f(x) = |x| − 1/2. projections for f(x) = −1/2, x < 0,<br />

f(x) = 1/2, x ≥ 0.<br />

<br />

,


100 Polynomial Approximation of Differential Equations<br />

Figure 6.2.1 shows the behavior of the approximating polynomials Πw,nf, when w<br />

is the Legendre weight function, n = 4, 6, 8, 10, and f is defined by f(x) = |x| − 1<br />

2 .<br />

A slower rate of convergence is measured at the point x = 0, where f presents a singu-<br />

lar<strong>it</strong>y in the derivative. In figure 6.2.2, a typical discontinuous function, i.e., f(x) =<br />

−1/2 if x < 0<br />

1/2 if x ≥ 0 , is approximated by Πw,nf, w<strong>it</strong>h w ≡ 1 and n = 7, 9, 11, 13. In s<strong>it</strong>uations<br />

like this one, the Gibbs phenomenon develops (see courant and hilbert(1953),<br />

Vol.1). As a consequence, the oscillations of the approximating polynomials, in the<br />

neighborhood of the point x = 0, stay bounded but do not damp for increasing n.<br />

Next, we evaluate the rate of convergence of the lim<strong>it</strong> in (6.2.8). This depends on<br />

the regular<strong>it</strong>y of the function f. The smoother the function, the faster is the convergence.<br />

This problem has been widely studied by many authors. The classical approach relates<br />

the speed of convergence to the so called modulus of continu<strong>it</strong>y of f (see timan(1963)).<br />

A more recent approach expresses the smoothness of f in terms of <strong>it</strong>s norm in appro-<br />

priate Sobolev spaces. We prefer the latter, since <strong>it</strong> is more su<strong>it</strong>able for the analysis of<br />

differential equations. Results in this direction are given for instance in babuˇska, sz-<br />

abo and katz(1981) for the Legendre case. Extensions are considered in canuto and<br />

quarteroni(1980) and canuto and quarteroni(1982a). Other references are listed<br />

in canuto, hussaini, quarteroni and zang(1988). The next theorem summarizes<br />

the results for Jacobi type approximations. Thus, {λj = j(j + α + β + 1)}j∈N denotes<br />

the sequence of eigenvalues corresponding to the Sturm-Liouville problem (1.3.1). We<br />

recall that a and w are related by the formula a(x) = (1 − x 2 )w(x), x ∈ I =] − 1,1[.<br />

According to defin<strong>it</strong>ion (5.7.1), f ∈ H k w(I) when f and <strong>it</strong>s derivatives of order up to k<br />

belong to L 2 w(I).<br />

Theorem 6.2.4 - Let k ∈ N, then there exists a constant C > 0 such that, for any<br />

f ∈ H k w(I), one has<br />

(6.2.15) f − Πw,nf L 2 w (I) ≤ C<br />

k 1<br />

(1<br />

2 k/2<br />

− x )<br />

n<br />

dkf dxk <br />

<br />

L 2 w (I)<br />

, ∀n > k.


Results in Approximation Theory 101<br />

Proof - For k = 0, (6.2.15) is a trivial consequence of (6.2.13). For k ≥ 1, let us set<br />

v := (1 − x 2 ) k w, x ∈ I, and observe that L 2 w(I) ⊂ L 2 v(I), since v ≤ w. This implies<br />

that f ′ ∈ L 2 w(I) ⊂ L 2 a(I) adm<strong>it</strong>s the representation<br />

(6.2.16) f ′ =<br />

∞<br />

cj u ′ j, a.e. in I,<br />

j=1<br />

where the cj’s are the Fourier coefficients of f. To show (6.2.16), we recall that {u ′ j }j≥1 ,<br />

up to normalizing factors, is an orthogonal basis of polynomials w<strong>it</strong>h respect to the<br />

weight function a (see (1.3.6) and (2.2.8)). Therefore, the Fourier coefficients dj, j ≥ 1,<br />

of f ′ w<strong>it</strong>h respect to this new basis, are<br />

(6.2.17) dj :=<br />

= λj<br />

<br />

<br />

I f ′ u ′ j<br />

a dx<br />

u ′ j 2 L 2 a (I)<br />

I fujw dx<br />

u ′ j 2 L 2 a (I)<br />

=<br />

<br />

= −<br />

<br />

I fujw dx<br />

uj 2 L 2 w (I)<br />

I f(u′ j a)′ dx<br />

u ′ j 2 L 2 a (I)<br />

= cj, j ≥ 1.<br />

Hence we get (6.2.16). In (6.2.17) we first integrated by parts (recalling that a(±1) = 0),<br />

then we used the fact that {uj}j∈N are the solutions of a Sturm-Liouville problem, and<br />

finally we used relation (2.2.15). For a correct justification of the equal<strong>it</strong>ies in (6.2.17),<br />

one should first argue w<strong>it</strong>h f ∈ C 1 ( Ī) and then approximate a general f ∈ H1 w(I) by a<br />

sequence of functions in C 1 ( Ī).<br />

In a similar way, we expand dk f<br />

dx k ∈ L 2 v(I) in series of the polynomial basis<br />

k<br />

d uj<br />

dxk <br />

,<br />

j≥k<br />

orthogonal w<strong>it</strong>h respect to the weight function v. The elements of this basis are solutions<br />

of a Sturm-Liouville problem w<strong>it</strong>h eigenvalues (j − k)(j + α + β + 1 + k), j ≥ k (see<br />

(1.3.6)). This yields<br />

(6.2.18)<br />

dkf =<br />

dxk ∞<br />

j=k<br />

cj<br />

dkuj , a.e. in I.<br />

dxk Finally, repeated application of (2.2.15) to successive derivatives of uj, j ≥ 1, gives for<br />

n > k:


102 Polynomial Approximation of Differential Equations<br />

(6.2.19) f − Πw,nf 2 L2 w (I) =<br />

≤<br />

<br />

=<br />

max<br />

j≥n+1<br />

≤<br />

∞<br />

j=n+1<br />

c 2 j<br />

k−1<br />

k−1<br />

i=0<br />

∞<br />

j=n+1<br />

c 2 j uj 2 L 2 w (I)<br />

<br />

[(j − i)(j + α + β + 1 + i)] −1<br />

<br />

<br />

dkuj dxk <br />

[(j − i)(j + α + β + 1 + i)] −1<br />

∞<br />

i=0<br />

k−1<br />

j=k<br />

<br />

[(n + 1 − i)(n + α + β + 2 + i)] −1<br />

<br />

<br />

dkf dxk <br />

<br />

i=0<br />

≤ C 2 n −2k<br />

<br />

<br />

(1 − x 2 ) k/2 dk f<br />

c 2 j<br />

dxk <br />

<br />

where C depends on α, β and k. Thus, we obtained (6.2.15).<br />

<br />

<br />

dk uj<br />

dx k<br />

2<br />

L 2 v (I)<br />

2<br />

L 2 w (I),<br />

<br />

<br />

2<br />

L2 v (I)<br />

<br />

<br />

2<br />

L2 v (I)<br />

The hypotheses of the previous theorem can be weakened by requiring only that f<br />

satisfies dm f<br />

dx m (1 − x 2 ) m/2 ∈ L 2 w(I), 0 ≤ m ≤ k.<br />

A proof based on similar arguments has been provided in bernardi and maday(1989)<br />

for ultraspherical polynomials. As pointed out in gottlieb and orszag(1977), p.33, a<br />

strict requirement to obtain estimates like (6.2.15), is that the polynomial basis consid-<br />

ered derives from a Sturm-Liouville problem, which is singular at the boundary points<br />

of the domain (see section 1.1).<br />

Equations (6.2.16) and (6.2.17) lead to the following remarkable relation, ∀n ≥ 1:<br />

(6.2.20) (Πw,nf) ′ = Πa,n−1f ′ , ∀f ∈ L 2 w(I) ∩ H 1 a(I).<br />

Inequal<strong>it</strong>y (6.2.15) is a typical example of a spectral error estimate. The rate of<br />

convergence depends only on the smoothness of the function f. In contrast to other<br />

approximation techniques, for instance those based on splines where the rate of con-<br />

vergence is controlled by the regular<strong>it</strong>y of the approximating functions in the domain,


Results in Approximation Theory 103<br />

the approximating polynomials, being analytic, do not affect the convergence behavior.<br />

In comparison w<strong>it</strong>h other methods, smaller values of the parameter n are in general<br />

sufficient to guarantee a satisfactory approximation error.<br />

Derivatives of the error f −Πw,nf can be also estimated, which result in a similar con-<br />

vergence behavior. Nevertheless, the rate of convergence is slower than that exhib<strong>it</strong>ed<br />

by other projection operators which are introduced later on. The reader interested in<br />

this field is referred to canuto and quarteroni(1982a), where other error estimates<br />

concerning Chebyshev and Legendre weights are taken into account. W<strong>it</strong>h techniques<br />

related to interpolation spaces (see section 5.7), formula (6.2.15) can be su<strong>it</strong>ably gener-<br />

alized to Sobolev spaces w<strong>it</strong>h real exponent. For other best approximation polynomials<br />

in more general functional spaces (for instance L p w(I) spaces, p > 1) the reader is ad-<br />

dressed to d<strong>it</strong>zian and totik(1987) and the references therein. Chebyshev expansions<br />

for functions w<strong>it</strong>h singular<strong>it</strong>ies at the endpoints of the interval [−1,1] are studied in<br />

boyd(1981).<br />

The same kind of result can be established for other systems of orthogonal poly-<br />

nomials. Laguerre polynomials require a l<strong>it</strong>tle care. Here λj = j, j ∈ N, and<br />

a(x) = xw(x) = x α+1 e −x , α > −1, x ∈ I =]0,+∞[. The inequal<strong>it</strong>y a(x) ≤ w(x) only<br />

holds for x ∈]0,1] ⊂ I, which implies that L 2 w(I) ⊂ L 2 a(I). Nevertheless, the proof of<br />

theorem 6.2.4 is still applicable and gives the next result.<br />

Theorem 6.2.5 - Let k ∈ N. Then there exists a constant C > 0 such that, for any<br />

f satisfying dm f<br />

dx m x m/2 ∈ L 2 w(I), 0 ≤ m ≤ k, one has<br />

(6.2.21) f − Πw,nf L 2 w (I) ≤ C<br />

k 1<br />

x<br />

k/2 √n<br />

dkf dxk <br />

<br />

L 2 w (I)<br />

, ∀n > k.<br />

Further estimates are given in maday, pernaud-thomas and vandeven(1985) for<br />

the case α = 0. In that paper, f is assumed to belong to the weighted space H k ω(I),<br />

w<strong>it</strong>h ω := e −(1−ǫ)x , ǫ > 0 arb<strong>it</strong>rarily small. In this s<strong>it</strong>uation, we obtain the inclusion<br />

H 1 ω(I) ⊂ L 2 w(I) ∩ H 1 a(I).


104 Polynomial Approximation of Differential Equations<br />

A proof identical to that of theorem 6.2.4 holds for Herm<strong>it</strong>e polynomials, where λj = 2j,<br />

j ∈ N, and a(x) = w(x) = e−x2, x ∈ R.<br />

Theorem 6.2.6 - Let k ∈ N. Then there exists a constant C > 0 such that, for any<br />

f ∈ H k w(R), one has<br />

(6.2.22) f − Πw,nf L 2 w (R) ≤ C<br />

In particular, (6.2.20) takes the form<br />

k 1<br />

<br />

d<br />

√n<br />

kf dxk <br />

<br />

L 2 w (R)<br />

(6.2.23) (Πw,nf) ′ = Πw,n−1f ′ , ∀f ∈ H 1 w(R), ∀n ≥ 1.<br />

, ∀n > k.<br />

From the examples illustrated here, <strong>it</strong> is clear that the speed of convergence to zero<br />

of the error is strictly related to the eigenvalues λj, j ∈ N (i.e., the spectrum of the<br />

Sturm-Liouville problem (1.1.1)). Somehow, this justifies the adoption of the adjective<br />

spectral, commonly used in the frame of these approximation methods.<br />

6.3 Inverse inequal<strong>it</strong>ies<br />

Inverse inequal<strong>it</strong>ies are an effective tool for the analysis of convergence in approximation<br />

theory. They are based on the fact that, in fin<strong>it</strong>e dimensional spaces, two given norms<br />

are always equivalent. Thus, we are allowed to give a bound to strong norms by using<br />

weaker norms. The drawback is that the constants in the equivalence relation, depend<br />

on the dimension of the space considered, so that <strong>it</strong> is not possible to extrapolate the<br />

same inequal<strong>it</strong>ies at the lim<strong>it</strong>. Classical examples in fin<strong>it</strong>e element method are given for<br />

instance in ciarlet(1978), p.140.<br />

Let us start w<strong>it</strong>h the Jacobi case (i.e., a(x) = (1 − x 2 )w(x), x ∈ I =] − 1,1[).


Results in Approximation Theory 105<br />

Lemma 6.3.1 - We can find a constant C > 0 such that, for any n ≥ 1<br />

(6.3.1) p L 2 w (I) ≤ Cn p L 2 a (I), ∀p ∈ Pn.<br />

Proof - Let m = n + 2. Since (1 − x 2 )p 2 ∈ P2m−2, formula (3.4.1) and theorem 3.4.1<br />

lead to<br />

(6.3.2)<br />

(6.3.3)<br />

<br />

I<br />

<br />

I<br />

p 2 a dx =<br />

p 2 w dx =<br />

m<br />

j=1<br />

m<br />

j=1<br />

p 2 (ξ (m)<br />

j ) w (m)<br />

j , ∀p ∈ Pn,<br />

p 2 (ξ (m)<br />

j )(1 − [ξ (m)<br />

j ] 2 ) w (m)<br />

j , ∀p ∈ Pn,<br />

where ξ (m)<br />

j , 1 ≤ j ≤ m, are the zeroes of P (α,β)<br />

m .<br />

Now, we can easily conclude the proof by noting that there exists a constant C ∗ > 0,<br />

depending only on α and β, such that, for any n ≥ 1<br />

(6.3.4) 1 − [ξ (m)<br />

j ] 2 ≥ C∗ C∗<br />

≥ 2<br />

m<br />

, 1 ≤ j ≤ m.<br />

9n2 In fact, from (1.3.2) and (3.1.17), the straight-line tangent to P (α,β)<br />

m<br />

at the point<br />

2(α+1)<br />

(α,β)<br />

x = 1 vanishes at ˆx := 1 − m(m+α+β+1) . It is easy to see that P m is convex in<br />

the interval [ξ (m)<br />

m ,1], hence ˆx is an upper bound for the zeroes of P (α,β)<br />

m . A similar<br />

argument is valid near the point x = −1 which proves (6.3.4).<br />

Relation (6.3.1) is an example of inverse inequal<strong>it</strong>y. Actually, the norm in L 2 w(I)<br />

is in general bigger than the norm in L 2 a(I), because w ≥ a, but the inequal<strong>it</strong>y in the<br />

oppos<strong>it</strong>e direction holds in the fin<strong>it</strong>e dimensional space Pn, for any fixed n ∈ N.<br />

More interesting cases are obtained for inverse inequal<strong>it</strong>ies that involve derivatives<br />

of polynomials. Examples have been given in section 2.5. Using Sobolev norms, we can<br />

establish similar results.<br />

Theorem 6.3.2 - We can find a constant C > 0 such that, for any n ≥ 1<br />

(6.3.5) p ′ L 2 a (I) ≤ Cn p L 2 w (I), ∀p ∈ Pn,


106 Polynomial Approximation of Differential Equations<br />

(6.3.6) p ′ L 2 w (I) ≤ Cn 2 p L 2 w (I), ∀p ∈ Pn.<br />

Proof - We set p = n<br />

k=0 ckuk, where {uk}k∈N is the sequence of Jacobi polynomials<br />

relative to the weight function w. From relations (2.2.15) and (2.2.8), we get<br />

(6.3.7) p ′ 2 L2 a (I) =<br />

=<br />

n<br />

k=1<br />

c 2 kλk uk 2 L2 w (I) ≤ λn<br />

n<br />

k=1<br />

n<br />

k=0<br />

c 2 k u ′ k 2 L 2 a (I)<br />

c 2 k uk 2 L 2 w (I) ≤ C2 n 2 p 2 L 2 w (I),<br />

where we observed that the eigenvalue λn behaves like n 2 . To obtain (6.3.6), we first<br />

apply lemma 6.3.1 to the polynomial p ′ and then use (6.3.5).<br />

As a byproduct of the previous theorem, <strong>it</strong> is not difficult to get, for any n ≥ 1<br />

(6.3.8) p H m w (I) ≤ Cn 2(m−k) p H k w (I), ∀p ∈ Pn, ∀m,k ∈ N, m ≥ k.<br />

When w is the Legendre weight function, relation (6.3.8) is the Schmidt inequal<strong>it</strong>y (see<br />

bellman(1944)). Using different arguments, (6.3.8) was established in canuto and<br />

quarteroni (1982a) for Chebyshev weights.<br />

We study now the Laguerre case (a(x) = xw(x) = x α+1 e −x , α > −1, x ∈ I =]0,+∞[).<br />

Lemma 6.3.3 - We can find two constants C1,C2 > 0 such that, for any n ≥ 1<br />

(6.3.9)<br />

C1<br />

√<br />

√ pL2 a (I) ≤ pL2 w (I) ≤ C2 n pL2 a (I), ∀p ∈ Pn.<br />

n<br />

Proof - As done in the proof of lemma 6.3.1, let ξ (m)<br />

j , 1 ≤ j ≤ m, be the zeroes of<br />

L (α)<br />

m , m = n + 1. Then one has the inequal<strong>it</strong>y C ∗ 1 1<br />

n<br />

≤ ξ(m)<br />

j<br />

≤ C∗ 2n, 1 ≤ j ≤ m (see<br />

section 3.1), where C ∗ 1,C ∗ 2 > 0 do not depend on n. Now we can easily deduce (6.3.9).


Results in Approximation Theory 107<br />

Using the same proof of theorem 6.3.2, we also deduce the the following propos<strong>it</strong>ion.<br />

Theorem 6.3.4 - We can find a constant C > 0 such that, for any n ≥ 1<br />

(6.3.10) p ′ L 2 a (I) ≤ C √ n p L 2 w (I), ∀p ∈ Pn,<br />

(6.3.11) p ′ L 2 w (I) ≤ Cn p L 2 w (I), ∀p ∈ Pn.<br />

Finally, we are left w<strong>it</strong>h the Herm<strong>it</strong>e case (i.e., a(x) = w(x) = e−x2, x ∈ R). Since<br />

we have the ident<strong>it</strong>y a ≡ w, lemma 6.3.1 is not useful. We also have the counterpart<br />

of theorem 6.3.2, which states:<br />

Theorem 6.3.5 - We can find a constant C > 0 such that, for any n ≥ 1<br />

(6.3.12) p ′ L 2 w (R) ≤ C √ n p L 2 w (R), ∀p ∈ Pn.<br />

Proof - We argue exactly as in (6.3.7).<br />

The problem of giving a bound to the norm in C0 ( Ī), by means of the norm in<br />

L 2 w(I), has arisen in section 3.9, where other inverse inequal<strong>it</strong>ies are presented.<br />

6.4 Other projection operators<br />

We focus our attention on theorem 6.2.2. As we can see from the next propos<strong>it</strong>ion, the<br />

expression (6.2.5) can actually be used to define the polynomial Πw,nf.


108 Polynomial Approximation of Differential Equations<br />

Theorem 6.4.1 - For a given f ∈ L 2 w(I), let ψ ∈ Pn such that<br />

(6.4.1) (f − ψ,φ) L 2 w (I) = 0, ∀φ ∈ Pn.<br />

Then we have ψ = Πw,nf.<br />

Proof - We can wr<strong>it</strong>e ψ = n<br />

k=0 dkuk for some coefficients dk, 0 ≤ k ≤ n. For any<br />

0 ≤ j ≤ n, choosing φ = uj as test function, (6.4.1) yields<br />

(6.4.2) (f,uj) L 2 w (I) =<br />

n<br />

k=0<br />

dk (uk,uj) L2 w (I) = dj uj 2 L2 w (I), 0 ≤ j ≤ n.<br />

Thus, by (6.2.2), the dk’s are the first n+1 Fourier coefficients of f. Hence ψ = Πw,nf.<br />

This result is the starting point for the defin<strong>it</strong>ion of new projection operators. For<br />

any m ∈ N and n ≥ m, we introduce the operator Π m w,n : H m w (I) → Pn, such that<br />

(6.4.3) (f − Π m w,nf,φ) H m w (I) = 0, ∀φ ∈ Pn.<br />

The inner product in H m w (I) has been defined in (5.6.2) and (5.7.2). Of course we<br />

have Π 0 w,n = Πw,n. One easily verifies that the polynomial Π m w,nf ∈ Pn is uniquely<br />

determined and represents the orthogonal projection of f in the H m w (I) norm. One<br />

checks that this projection satisfies a minimization problem like (6.2.4), namely<br />

(6.4.4) f − Π m w,nfH m<br />

w (I) = inf f − ψHm w (I).<br />

ψ∈Pn<br />

Moreover, we have for n ≥ m<br />

(6.4.5) Π m w,nf H m w (I) ≤ f H m w (I), ∀f ∈ H m w (I).<br />

As before, the expression in (6.4.4) tends to zero when n → +∞, w<strong>it</strong>h a rate depending<br />

on the degree of smoothness of f. For simplic<strong>it</strong>y, we restrict the analysis to the case<br />

m = 1 and we just consider the Jacobi weight w(x) = (1−x) α (1+x) β , x ∈ I =]−1,1[.


Results in Approximation Theory 109<br />

Theorem 6.4.2 - Let −1 < α < 1, −1 < β < 1 and k ≥ 1, then we can find a<br />

constant C > 0 such that, for any f ∈ H k w(I), one has<br />

(6.4.6) f − Π 1 w,nf H 1 w (I) ≤ C<br />

k−1 <br />

1 <br />

(1 − x<br />

n<br />

2 ) (k−1)/2 dkf dxk <br />

<br />

<br />

Proof - We first consider f ∈ C ∞ ( Ī). Let x0 ∈ I and set<br />

(6.4.7) qn(x) :=<br />

From (6.4.4) we can wr<strong>it</strong>e<br />

<br />

f(x0) +<br />

x<br />

x0<br />

(Πw,n−1f ′ <br />

)(t) dt<br />

(6.4.8) f − Π 1 w,nf H 1 w (I) ≤ f − qn H 1 w (I)<br />

=<br />

L 2 w (I)<br />

<br />

f − qn 2 L2 w (I) + f ′ − Πw,n−1f ′ 2 L2 w (I)<br />

1<br />

2<br />

.<br />

On the other hand, we have by the Schwarz inequal<strong>it</strong>y<br />

(6.4.9) f − qn 2 L2 w (I) =<br />

<br />

≤<br />

I<br />

<br />

=<br />

<br />

I<br />

x<br />

x0<br />

<br />

I<br />

<br />

f(x) − f(x0) −<br />

x<br />

(f ′ − Πw,n−1f ′ )(t) w(t)<br />

x<br />

(f ′ − Πw,n−1f ′ ) 2 (t) w(t) dt<br />

x0<br />

≤ f ′ − Πw,n−1f ′ 2 L 2 w (I)<br />

<br />

I<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

·<br />

<br />

<br />

<br />

<br />

x<br />

x0<br />

x0<br />

x<br />

x0<br />

∈ Pn, n ≥ 1.<br />

, ∀n > k.<br />

(Πw,n−1f ′ 2 )(t)dt w(x)dx<br />

2 dt<br />

w(x) dx<br />

w(t)<br />

w −1 <br />

<br />

(t) dt<br />

w(x) dx<br />

w −1 <br />

<br />

(t)dt<br />

w(x) dx.<br />

We note that the last integral in (6.4.9) is fin<strong>it</strong>e when −1 < α < 1 and −1 < β < 1.<br />

At this point we can conclude after subst<strong>it</strong>uting (6.4.9) in (6.4.8) and recalling theorem<br />

6.2.4.


110 Polynomial Approximation of Differential Equations<br />

The proof is completed by approximating f ∈ H k w(I) w<strong>it</strong>h a sequence of functions in<br />

C∞ ( Ī) (see section 5.7).<br />

The proof of the above statement was formerly given in maday and quarteroni(1981)<br />

for Legendre and Chebyshev weights. The extension to other ultraspherical weights is<br />

provided in bernardi and maday(1989). In those papers, one also finds estimates of<br />

the same error in other norms. For instance, when −1 < α = β < 1 and k ≥ 1, there<br />

exists a constant C > 0 such that, for any f ∈ H k w(I)<br />

(6.4.10) f − Π 1 w,nf L 2 w (I) ≤ C<br />

k 1<br />

fH k<br />

w n<br />

(I), ∀n > k.<br />

Su<strong>it</strong>able projections can be also defined in H 1 0,w(I) ⊂ H 1 w(I) (see defin<strong>it</strong>ion (5.7.6)).<br />

First of all we introduce the space<br />

(6.4.11) P 0 n :=<br />

<br />

p ∈ Pn<br />

<br />

<br />

p(±1) = 0 , n ≥ 2.<br />

Assuming that −1 < α < 1 and −1 < β < 1, we examine two projection operators.<br />

For any n ≥ 2, we consider Π 1 0,w,n : H 1 0,w(I) → P 0 n and ˆ Π 1 0,w,n : H 1 0,w(I) → P 0 n such<br />

that, for f ∈ H 1 0,w(I), we have<br />

(6.4.12)<br />

(6.4.13)<br />

<br />

<br />

(f − Π<br />

I<br />

1 0,w,nf) ′ φ ′ w dx = 0, ∀φ ∈ P 0 n,<br />

(f −<br />

I<br />

ˆ Π 1 0,w,nf) ′ (φw) ′ dx = 0, ∀φ ∈ P 0 n.<br />

Existence and uniqueness for Π 1 0,w,nf are obtained by recalling defin<strong>it</strong>ion (5.7.7) and<br />

arguing as in theorem 6.2.1. Furthermore, <strong>it</strong> is easy to verify<br />

(6.4.14) f − Π 1 0,w,nfH 1<br />

0,w (I) = inf<br />

ψ∈P0 f − ψH1 0,w (I).<br />

n<br />

Error estimates for Π 1 0,w,n can be inferred from estimates of other operators, as illus-<br />

trated by the following propos<strong>it</strong>ion.


Results in Approximation Theory 111<br />

Theorem 6.4.3 - There exists a constant C > 0 such that, for any f ∈ H 1 0,w(I), we<br />

have<br />

(6.4.15) f − Π 1 0,w,nf H 1 0,w (I) ≤ C f − Π 1 w,nf H 1 w (I), ∀n ≥ 2.<br />

Proof - We define<br />

(6.4.16) ψ(x) := (Π 1 w,nf)(x) + 1<br />

2<br />

<br />

(1 + x)(f − Π 1 w,nf)(1)<br />

+ (1 − x)(f − Π 1 <br />

w,nf)(−1) , x ∈ I.<br />

Recalling that f(±1) = 0, we have ψ ∈ P 0 n. Hence, subst<strong>it</strong>uting in (6.4.14), we get<br />

(6.4.17) f − Π 1 0,w,nf H 1 0,w (I) ≤ f − Π 1 w,nf H 1 0,w (I)<br />

+ 1<br />

2<br />

<br />

<br />

d<br />

dx<br />

<br />

(1 + x)(f − Π 1 w,nf)(1) + (1 − x)(f − Π 1 <br />

L<br />

w,nf)(−1)<br />

2<br />

w (I)<br />

≤ f − Π 1 w,nf H 1 w (I) + C ∗ f − Π 1 w,nf C 0 (Ī) ,<br />

where C ∗ > 0 is a constant. Now, recalling (5.7.4), we immediately conclude the proof.<br />

The analysis of the operator ˆ Π 1 0,w,n is more involved and will be considered in<br />

section 9.4. Of course, when w ≡ 1 is the Legendre weight function, the relations<br />

(6.4.12) and (6.4.13) are identical.<br />

6.5 Convergence of the Gaussian formulas<br />

High order integration formulas were introduced in sections 3.4, 3.5 and 3.6. Here we<br />

consider their asymptotic properties for increasing number of nodes. In the case of<br />

Jacobi weights, we begin w<strong>it</strong>h the corollary of a convergence result due to Stekloff and<br />

Fejér (see szegö(1939), p.342).


112 Polynomial Approximation of Differential Equations<br />

Theorem 6.5.1 - Let f be such that fw is Riemann integrable in I =] − 1,1[. Then<br />

we have<br />

(6.5.1) lim<br />

n→+∞<br />

n<br />

j=1<br />

f(ξ (n)<br />

j ) w (n)<br />

j<br />

<br />

= lim Iw,nf w dx =<br />

n→+∞<br />

I<br />

<br />

I<br />

fw dx.<br />

Proof - We show a simplified proof for f ∈ C0 ( Ī). The general case is studied in<br />

stekloff(1916). We use the polynomial of best uniform approximation of f (see section<br />

6.1). For any n ≥ 1, taking into account that [Ψ∞,n−1(f) − Iw,nf] ∈ Pn−1, we have<br />

(6.5.2)<br />

<br />

<br />

<br />

<br />

<br />

(f −Iw,nf)wdx <br />

≤<br />

<br />

<br />

<br />

<br />

<br />

(f −Ψ∞,n−1(f))wdx <br />

+<br />

<br />

<br />

<br />

<br />

I<br />

≤ f −Ψ∞,n−1(f) C 0 (Ī)<br />

I<br />

⎛<br />

<br />

⎝<br />

I<br />

w dx +<br />

n<br />

j=1<br />

w (n)<br />

j<br />

⎞<br />

n<br />

(Ψ∞,n−1(f) −Iw,nf)(ξ<br />

j=1<br />

(n)<br />

j ) w (n)<br />

j<br />

⎠ = 2 f −Ψ∞,n−1(f) C 0 (Ī)<br />

<br />

I<br />

<br />

<br />

<br />

<br />

w dx,<br />

where we made use of Iw,nf(ξ (n)<br />

j ) = f(ξ (n)<br />

j ), 1 ≤ j ≤ n. Now, because of (6.1.5), the<br />

last term in (6.5.2) tends to zero. Moreover, when f is smooth, we can establish the<br />

rate of convergence using (6.1.7).<br />

For Laguerre and Herm<strong>it</strong>e quadrature formulas we can state similar propos<strong>it</strong>ions (see<br />

uspensky(1928) and davis and rabinow<strong>it</strong>z(1984)).<br />

Theorem 6.5.2 - Let f be such that fw is Riemann integrable in I =]0,+∞[, where<br />

w is the Laguerre weight function. Assume the existence of x0 ∈ I and ǫ > 0 such that<br />

|f(x)w(x)| ≤ x −1−ǫ , ∀x > x0. Under these assumptions (6.5.1) holds.<br />

Theorem 6.5.3 - Let f be such that fw is Riemann integrable in I = R, where w<br />

is the Herm<strong>it</strong>e weight function. Assume the existence of x0 ∈ I and ǫ > 0 such that<br />

|f(x)w(x)| ≤ x −1−ǫ , ∀|x| > x0. Under these assumptions (6.5.1) holds.


Results in Approximation Theory 113<br />

The rate of convergence in (6.5.1) is measured by the degree of smoothness of the<br />

function f. If f fulfills certain regular<strong>it</strong>y requirements, we can analytically express the<br />

error between the integral and the quadrature. The proof of the following theorem does<br />

not present difficulties and is given for instance in todd(1963).<br />

Theorem 6.5.4 - Let n ≥ 1 and f ∈ C2n ( Ī). Let us define<br />

(6.5.3)<br />

<br />

En(f) := fw dx −<br />

n<br />

I<br />

j=1<br />

f(ξ (n)<br />

j ) w (n)<br />

j .<br />

Then we can find ξ ∈ I such that<br />

<br />

2<br />

(6.5.4) (Jacobi) En(f) =<br />

2n+α+β+1 n! Γ(n + α + 1) Γ(n + β + 1)<br />

(2n)! Γ(2n + α + β + 1)<br />

× Γ(n + α + β + 1)<br />

<br />

Γ(2n + α + β + 2)<br />

(6.5.5) (Legendre) En(f) =<br />

(6.5.6) (Chebyshev) En(f) =<br />

(6.5.7) (Laguerre) En(f) =<br />

d2nf (ξ), α > −1, β > −1,<br />

dx2n 2 2n+1 [n!] 4<br />

(2n + 1) [(2n)!] 3<br />

π<br />

2 2n−1 (2n)!<br />

n! Γ(n + α + 1)<br />

(2n)!<br />

(6.5.8) (Herm<strong>it</strong>e) En(f) = n! √ π<br />

2 n (2n)!<br />

d2nf (ξ),<br />

dx2n d2nf (ξ),<br />

dx2n d2nf (ξ), α > −1,<br />

dx2n d2nf (ξ).<br />

dx2n Concerning Gauss-Lobatto formulas, we can recover a convergence result for f ∈<br />

C0 ( Ī), I =] − 1,1[, w<strong>it</strong>h the same proof given for theorem 6.5.1. A more general<br />

theorem, based on weaker assumptions on f, also holds in the Jacobi case:<br />

Theorem 6.5.5 - Let f : Ī → R, such that fw is Riemann integrable. Then we have<br />

(6.5.9)<br />

n<br />

lim f(η<br />

n→+∞<br />

(n)<br />

j ) ˜w (n)<br />

j<br />

<br />

= lim<br />

n→+∞<br />

<br />

Ĩw,nf w dx = fw dx.<br />

j=0<br />

I<br />

I


114 Polynomial Approximation of Differential Equations<br />

Proof - We first recall the equal<strong>it</strong>y<br />

(6.5.10) Ĩw,nf = [Ia,n−1(fq −1 )] q + f(−1) ˜ l (n)<br />

0 + f(1) ˜ l (n)<br />

n , n ≥ 1,<br />

where q(x) := (1 − x 2 ), x ∈ I, a = qw, and the Lagrange polynomials are defined<br />

in (3.2.8). To check (6.5.10) <strong>it</strong> is sufficient to note that the relation holds at the points<br />

η (n)<br />

j , 0 ≤ j ≤ n. Actually, due to (1.3.6), we have P (α+1,β+1)<br />

n−1 (η (n)<br />

j ) = 0, 1 ≤ j ≤ n −1.<br />

Hence, the η (n)<br />

j ’s are the nodes of a Gauss formula where α and β are respectively<br />

(n)<br />

replaced by α + 1 and β + 1. Therefore ( Ĩw,ng)(η j ) = (Ia,n−1g)(η (n)<br />

j ) = g(η (n)<br />

j ),<br />

1 ≤ j ≤ n − 1, for any function g. The verification of (6.5.10) at the points x = ±1 is<br />

trivial. Multiplying both sides of (6.5.10) by w and integrating in I, one gets<br />

(6.5.11) lim<br />

n→+∞<br />

<br />

[Ia,n−1(fq<br />

I<br />

−1 )] qw dx =<br />

(6.5.12) lim<br />

n→+∞<br />

<br />

I<br />

˜(n) l j w dx = lim<br />

n→+∞ ˜w(n) j<br />

<br />

I<br />

fq −1 a dx =<br />

<br />

I<br />

fw dx,<br />

= 0, j = 0 or j = n.<br />

The first lim<strong>it</strong> is a consequence of theorem 6.5.1, and (6.5.12) is obtained by studying<br />

the behavior for n → +∞ of the weights in (3.5.2).<br />

A statement like that of theorem 6.5.4 also exists for Gauss-Lobatto and Gauss-<br />

Radau formulas. We mention davis and rabinow<strong>it</strong>z(1984) for the treatment of the<br />

Legendre case and ghizzetti and ossicini(1970) for a more general discussion. Esti-<br />

mates in the Legendre case are also analyzed in d<strong>it</strong>zian and totik(1987), p.87.<br />

In the case of Laguerre Gauss-Radau formulas (see (3.6.1)), we obtain a propos<strong>it</strong>ion like<br />

theorem 6.5.2. This time the proof is based on the following relation:<br />

(6.5.13) Ĩw,nf = [Ia,n−1(fq −1 )] q + f(0) ˜l (n)<br />

0 , n ≥ 1,<br />

where q(x) = x, x ∈]0,+∞[, a = qw, and ˜ l (n)<br />

0 is given in (3.2.11).


Results in Approximation Theory 115<br />

6.6 Estimates for the interpolation operator<br />

The interpolation operators Iw,n and Ĩw,n have been introduced in section 3.3. In<br />

this section, we are concerned w<strong>it</strong>h estimating the rate of convergence to zero of the<br />

error between a function and <strong>it</strong>s interpolant. We first study the case of ultraspherical<br />

polynomials (ν := α = β).<br />

Theorem 6.6.1 - Let k ≥ 1 and −1 < ν ≤ 0. Then we can find a constant C > 0<br />

such that, for any f ∈ H k w(I), we have<br />

(6.6.1) f − Iw,nf L 2 w (I) ≤ C<br />

The same relation holds for the error f − Ĩw,nf.<br />

k−1/2 1<br />

(1<br />

2 −ν/2<br />

− x )<br />

n<br />

dkf dxk <br />

<br />

Proof - We follow the proof of theorem 6.5.1. One has<br />

(6.6.2) f − Iw,nf L 2 w (I) ≤ f − Ψ∞,n−1(f) L 2 w (I)<br />

+<br />

⎛<br />

⎝<br />

n<br />

(Ψ∞,n−1(f) − Iw,nf)<br />

j=1<br />

2 (ξ (n)<br />

j ) w (n)<br />

j<br />

⎞<br />

⎠<br />

1<br />

2<br />

L 2 w (I)<br />

≤ 2 f −Ψ∞,n−1(f) C 0 (Ī)<br />

Thus, recalling (6.1.7), we can find a constant C ∗ > 0 such that<br />

(6.6.3) f − Iw,nf L 2 w (I) ≤ C ∗<br />

≤ C ∗<br />

= C ∗<br />

k−1 1<br />

n<br />

k−1 1<br />

n<br />

sup<br />

|x1−x2|


116 Polynomial Approximation of Differential Equations<br />

where we used the Schwarz inequal<strong>it</strong>y. This proves the statement. Concerning the error<br />

f − Ĩw,nf, we apply the same arguments. This time, the quadrature formula is not<br />

exact for polynomials in P2n. However, we obtain an inequal<strong>it</strong>y like (6.6.2), thanks to<br />

theorem 3.8.2.<br />

For the Chebyshev case (ν = −1/2), a sharper estimate of the error is readily available<br />

(6.6.4) f − Iw,nf L 2 w (I) ≤ C<br />

k 1<br />

n<br />

f H k w (I),<br />

where k ≥ 1 and w(x) = 1/ √ 1 − x 2 , x ∈ I. The proof of this inequal<strong>it</strong>y relies on<br />

the possibil<strong>it</strong>y of using results in approximation theory for trigonometric polynomials<br />

(see jackson(1930) or zygmund(1988)), related to Chebyshev polynomials via (1.5.6).<br />

The same procedure applies for the interpolation operator corresponding to Chebyshev<br />

Gauss-Lobatto nodes.<br />

To obtain convergence results for Iw,n when ν > 0, we recall the expression<br />

(6.6.5) Iw,nf = [ Ĩv,n+1(fq)] q −1 , ∀n ∈ N,<br />

where q(x) := (1 − x 2 ), x ∈ I, and v := q −1 w. Actually, both the terms in (6.6.5)<br />

are polynomials in Pn−1, coinciding at the nodes ξ (n)<br />

j , 1 ≤ j ≤ n, as a consequence of<br />

relation (1.3.6). Therefore, one gets<br />

(6.6.6) f − Iw,nf L 2 w (I) = (fq) − Ĩv,n+1(fq) L 2 v (I).<br />

Thus, an estimate for 0 < ν ≤ 1, can be derived from theorem 6.6.1, by noting that v<br />

is an ultraspherical weight function, whose exponent is negative.<br />

Convergence results for Ĩw,n in the case −1 < ν < 1 are presented in bernardi and<br />

maday(1989). Proofs are based on techniques borrowed from the theory of interpolation<br />

of Sobolev spaces (see section 5.7).


Results in Approximation Theory 117<br />

We observe that the inequal<strong>it</strong>y (6.6.2) is satisfied for all the Jacobi weights. There-<br />

fore, by (6.1.7), the rate of convergence can be estimated in the general case. For<br />

instance, for any f ∈ Ck ( Ī), k ≥ 1, the following inequal<strong>it</strong>y holds:<br />

(6.6.7) f − Iw,nf L 2 w (I) ≤ C<br />

k 1<br />

<br />

d<br />

n<br />

kf dxk <br />

<br />

C 0 (Ī)<br />

, ∀n ≥ k,<br />

where w(x) = (1 − x) α (1 + x) β , α > −1, β > −1, x ∈ I. The same statement applies<br />

for the error f − Ĩw,nf. We expect that (6.6.7) can be further refined by weakening the<br />

assumptions on f and, w<strong>it</strong>h the same rate of convergence, replacing the norm on the<br />

right-hand side of (6.6.7) w<strong>it</strong>h a weaker one (see also section 6.8).<br />

By means of inverse inequal<strong>it</strong>ies, we also give a bound to the derivative of the error<br />

as follows:<br />

(6.6.8) (f − Iw,nf) ′ L 2 w (I) ≤ (f − Π 1 w,nf) ′ L 2 w (I) + (Π 1 w,nf − Iw,nf) ′ L 2 w (I)<br />

≤ f − Π 1 w,nf H 1 w (I) + Cn 2 Π 1 w,nf − Iw,nf L 2 w (I),<br />

where Π 1 w,n is defined in (6.4.3) and where we used the inequal<strong>it</strong>y (6.3.6) for the<br />

polynomial p := Π 1 w,nf − Iw,nf ∈ Pn. Assuming a sufficient regular<strong>it</strong>y for f, the last<br />

terms in (6.6.8) tend to zero for n → +∞.<br />

It is clear now that the aliasing errors Aw,nf and Ãw,nf, introduced in section<br />

4.2, decay to zero for n → +∞, w<strong>it</strong>h a rate depending on the smoothness of f. From<br />

(6.2.4) we know that<br />

(6.6.9) f − Πw,n−1fL2 w (I) ≤ f − Iw,nfL2 w (I), ∀f ∈ C 0 ( Ī), ∀n ≥ 1.<br />

We can be more precise about this estimate.<br />

Theorem 6.6.2 - For any n ≥ 1 and any f ∈ C0 ( Ī), we have<br />

(6.6.10) f − Iw,nf 2 L 2 w (I) = f − Πw,n−1f 2 L 2 w (I) + Aw,nf 2 L 2 w (I),<br />

(6.6.11) f − Ĩw,nf 2 L 2 w (I) = f − Πw,nf 2 L 2 w (I) + Ãw,nf 2 L 2 w (I).


118 Polynomial Approximation of Differential Equations<br />

Proof - Expand all the squared norms in (6.6.10) (or (6.6.11)) in terms of the inner<br />

product in L 2 w(I) and recall (6.2.5).<br />

Another characterization of the aliasing error is given by (see (3.3.1) and (6.2.11))<br />

<br />

∞<br />

<br />

∞<br />

=<br />

(6.6.12) Aw,nf = Iw,n(f − Πw,n−1f) = Iw,n<br />

k=n<br />

ckuk<br />

k=n+1<br />

ck(Iw,nuk),<br />

where ck, k ∈ N, are the Fourier coefficients of f and uk = P (α,β)<br />

k , k ∈ N. The<br />

last equal<strong>it</strong>y in (6.6.12) follows by observing that, for any n ≥ 1, Iw,n is a continuous<br />

operator in C 0 ( Ī) (see section 6.8). A similar relation holds for Ãw,nf. In the Chebyshev<br />

case, the formula (6.6.12) takes a simplified form, thanks to the periodic<strong>it</strong>y of the cosine<br />

function, i.e.<br />

(6.6.13) Aw,nf =<br />

n−1 <br />

∞<br />

m=1 j=1<br />

(c2jn+m − c2jn−m)Tm.<br />

Moreover, we note that Aw,nTjn ≡ 0, ∀j ≥ 1. A discussion of the effects of the aliasing<br />

error in the computation of solutions of boundary value problems are given in canuto,<br />

hussaini, quarteroni and zang(1988).<br />

We analyse now the case when I is not bounded. Let us start w<strong>it</strong>h the Laguerre<br />

approximations (w(x) = x α e −x , α > −1, x ∈ I =]0,+∞[).<br />

Theorem 6.6.3 - Let δ ∈]0,1[ and let f ∈ C 0 ( Ī) satisfy limx→+∞ f(x)e −δx = 0.<br />

Then we have<br />

(6.6.14) lim<br />

n→+∞ f − Iw,nf L 2 w (I) = 0.<br />

The same relation holds for the error f − Ĩw,nf, relative to the Gauss-Radau interpo-<br />

lation operator.<br />

Proof - For any pn ∈ Pn−1, we have (compare w<strong>it</strong>h (6.6.2))<br />

(6.6.15) f − Iw,nf L 2 w (I) ≤ f − pn L 2 w (I) +<br />

⎛<br />

⎝<br />

n<br />

(pn − f)<br />

j=1<br />

2 (ξ (n)<br />

j ) w (n)<br />

j<br />

⎞<br />

⎠<br />

1<br />

2<br />


Results in Approximation Theory 119<br />

≤ sup<br />

x∈Ī <br />

|f − pn| √ <br />

v (x)<br />

<br />

I<br />

v −1 1<br />

2<br />

w dx<br />

⎛<br />

n<br />

+ ⎝<br />

v<br />

j=1<br />

−1 (ξ (n)<br />

j ) w (n)<br />

j<br />

⎞ 1<br />

2 <br />

where v(x) := e−δx , x ∈ Ī. Due to theorem 6.5.2, the summation on the right hand<br />

side of (6.6.15) converges to <br />

I v−1 wdx < +∞. Hence, <strong>it</strong> is bounded by a constant<br />

independent of n. According to theorem 6.1.4, we can choose the sequence {pn}n≥1 to<br />

obtain (6.6.14).<br />

On the determination of the rate of convergence, we just mention a result given in<br />

maday, pernaud-thomas and vandeven(1985), for the case α = 0 (w(x) = e −x ).<br />

Theorem 6.6.4 - Let k ≥ 1 and δ ∈]0,1[. Then we can find a constant C > 0 such<br />

that, for any f ∈ H k v (I), w<strong>it</strong>h v(x) := e −δx , one has<br />

(6.6.16) f − Iw,nf L 2 w (I) ≤ C<br />

k−1 1<br />

√n<br />

The same relation holds for the error f − Ĩw,nf.<br />

⎠<br />

f H k v (I), ∀n ≥ 1.<br />

We argue similarly for the Herm<strong>it</strong>e weight function (w(x) = e−x2). A theorem like 6.6.3<br />

is derived from theorems 6.1.5 and 6.5.3. Further results are considered in the next<br />

section.<br />

6.7 Laguerre and Herm<strong>it</strong>e functions<br />

In the analysis of differential equations in unbounded domains, the solution will be<br />

required to decay at infin<strong>it</strong>y w<strong>it</strong>h an exponential rate. Therefore, approximation by a<br />

,


120 Polynomial Approximation of Differential Equations<br />

polynomial cannot give accurate results (see gottlieb and orszag(1977), p.45). It is<br />

thus necessary to introduce new families of approximating functions. We set v(x) :=<br />

x α e x , α > −1, x ∈ I. For n ∈ N, we define Sn to be the space of the functions (named<br />

Laguerre functions) of the form pe −x , where p ∈ Pn. By virtue of (2.2.6), this is an<br />

orthogonal set of functions where the inner product is weighted by v. We present a<br />

preliminary result.<br />

Lemma 6.7.1 - Let f = ge x then, for any k ∈ N, we have<br />

(6.7.1) x m/2 dm f<br />

dx m ∈ L2 w(I), 0 ≤ m ≤ k ⇐⇒ x k/2 dk g<br />

dx k ∈ L2 v(I).<br />

Proof - This is a direct consequence of the formula<br />

(6.7.2)<br />

=<br />

k<br />

m=0<br />

<br />

I<br />

<br />

k Γ(α + k + 1)<br />

m Γ(α + m + 1)<br />

k d g<br />

dxk 2<br />

x α+k e x dx<br />

which is proven by induction (see funaro(1991)).<br />

<br />

I<br />

m d f<br />

dxm 2 x α+m e −x dx,<br />

The operator Π ∗ v,n : L 2 v(I) → Sn, n ∈ N, is defined as follows:<br />

(6.7.3) Π ∗ v,ng := [Πw,n(ge x )]e −x , ∀g ∈ L 2 v(I).<br />

Combining theorem 6.2.5 and relation (6.7.2), we get the following propos<strong>it</strong>ion.<br />

Theorem 6.7.2 - Let k ∈ N. Then there exists a constant C > 0 such that, for any g<br />

satisfying dk g<br />

dx k x k/2 ∈ L 2 v(I), one has<br />

(6.7.4) g − Π ∗ v,ng L 2 v (I) ≤ C<br />

k 1<br />

x<br />

k/2 √n<br />

dkg dxk <br />

<br />

L 2 v (I)<br />

, ∀n > k.


Results in Approximation Theory 121<br />

Furthermore, we have the following inverse inequal<strong>it</strong>y.<br />

Theorem 6.7.3 - We can find a constant C > 0 such that, for any n ≥ 1<br />

(6.7.5) q ′ L 2 v (I) ≤ Cn q L 2 v (I), ∀q ∈ Sn.<br />

Proof - Set q = pe −x , p ∈ Pn, and recall (6.3.11).<br />

Next, we define the operator I ∗ v,n : C 0 ( Ī) → Sn−1, n ≥ 1, such that<br />

(6.7.6) I ∗ v,ng := [Iw,n(ge x )]e −x , ∀g ∈ C 0 ( Ī).<br />

In a similar way, one defines the interpolation operator Ĩ∗ v,n, based on the Gauss-Radau<br />

nodes. On the basis of theorem 6.6.4, one recovers estimates for the error g − I ∗ v,ng .<br />

Now let us set v(x) := w−1 (x) = ex2, x ∈ R, and define for any n ∈ N, the space Sn of<br />

the so called Herm<strong>it</strong>e functions. Any element of Sn is of the form pw, where p ∈ Pn.<br />

Furthermore, we introduce the operator Π ∗ v,n : L 2 v(R) → Sn, n ∈ N, such that<br />

(6.7.7) Π ∗ v,ng := [Πw,n(gv)]w, ∀g ∈ L 2 v(R).<br />

As in the previous case, we begin w<strong>it</strong>h a basic result.<br />

Lemma 6.7.4 - The following implication: f ∈ H k w(R) ⇔ g := fw ∈ H k v (R), is<br />

satisfied for any k ∈ N. Moreover, the corresponding norms are equivalent.<br />

Proof - It is possible to find real coefficients γ (k)<br />

m , 0 ≤ k ≤ m (see funaro(1991)),<br />

such that<br />

(6.7.8)<br />

<br />

R<br />

k d g<br />

dxk 2<br />

v dx =<br />

k<br />

m=0<br />

γ (k)<br />

m<br />

This completes the proof by virtue of (5.5.10) and (5.6.4).<br />

<br />

R<br />

m d f<br />

dxm 2 w dx.


122 Polynomial Approximation of Differential Equations<br />

Taking into account lemma 6.7.4, theorems 6.2.6 and 6.3.5 are respectively generalized<br />

to the case of Herm<strong>it</strong>e functions as follows.<br />

Theorem 6.7.5 - Let k ∈ N, then there exists a constant C > 0 such that, for any<br />

g ∈ H k v (R), one has<br />

(6.7.9) g − Π ∗ v,ng L 2 v (R) ≤ C<br />

k 1<br />

√n<br />

g H k v (R), ∀n > k.<br />

Theorem 6.7.6 - We can find a constant C > 0 such that, for any n ≥ 1<br />

(6.7.10) q ′ L 2 w (R) ≤ C √ n q L 2 w (R), ∀q ∈ Sn.<br />

We finally define I ∗ v,n : C 0 (R) → Sn−1, n ≥ 1, such that<br />

(6.7.11) I ∗ v,n := [Iw,n(gv)]w, ∀g ∈ C 0 (R).<br />

A theoretical convergence analysis for the operator I ∗ v,n for n → +∞ is considered in<br />

funaro and kavian(1988). Here we state the main result (we recall that the Sobolev<br />

spaces w<strong>it</strong>h real exponent have been introduced in section 5.6).<br />

Theorem 6.7.7 - Let ǫ > 0 and s ≥ 1 + ǫ. Then there exists a constant C > 0 such<br />

that, for any g ∈ H s v(R), we have<br />

(6.7.12) g − I ∗ v,ng L 2 v (R) ≤ C<br />

s−1−ǫ 1<br />

√n gHs v (R), ∀n ≥ 1.<br />

Other results, concerning the rate of convergence of the so called normalized Herm<strong>it</strong>e<br />

functions, are provided in boyd(1984).


Results in Approximation Theory 123<br />

6.8 Refinements<br />

The aim of the previous sections was to present an overview of the fundamental results<br />

on the approximation of functions by orthogonal polynomials. Further developments can<br />

be taken into account. In general, improvements require a certain skill in dealing w<strong>it</strong>h<br />

functional spaces and their properties. We therefore confine ourselves to the discussion<br />

of some basic facts.<br />

In the book of timan(1963), p.262, the inequal<strong>it</strong>y (6.1.7) is generalized as follows:<br />

(6.8.1) |f − Ψ∞,n(f)|(x) ≤ C [hn(x)] k<br />

where hn(x) := 1<br />

n<br />

sup<br />

|x1−x2|


124 Polynomial Approximation of Differential Equations<br />

w<strong>it</strong>h different techniques, shows a faster convergence behavior. Although the inequal<strong>it</strong>y<br />

(6.6.7) ensures the same kind of convergence, <strong>it</strong> requires f to be more regular. In<br />

nevai(1976) numerous extensions are provided, and in maday(1991) an estimate like<br />

(6.6.4) is proven for Legendre Gauss-Lobatto points.<br />

For analytic functions, the error estimates for the interpolation operator are expected<br />

to exhib<strong>it</strong> an exponential decay. Results in this direction are given in tadmor(1986)<br />

for Chebyshev expansions. Various improvements and estimates in different weighted<br />

spaces are considered in d<strong>it</strong>zian and totik(1987).<br />

To give an idea of the difficulties one may face when working w<strong>it</strong>h the interpolation<br />

operators, we note for instance that an inequal<strong>it</strong>y like (6.2.7) is not verified, when<br />

replacing Πw,n by Iw,n. This shows that the operator Iw,n is not continuous in<br />

the L2 w(I) norm, even if f ∈ C0 ( Ī). To check this fact, we construct a sequence of<br />

functions fk ∈ C0 (n)<br />

( Ī), k ∈ N, such that fk(ξ j ) = 1, 1 ≤ j ≤ n, ∀k ∈ N. Therefore,<br />

one has Iw,nfk ≡ 1, ∀k ∈ N, which implies that Iw,nfkL2 w (I) = 0 is constant w<strong>it</strong>h<br />

respect to k. On the other hand, every fk can be defined at the remaining points in<br />

such a way that limk→+∞ fk L 2 w (I) = 0. This is in contrast w<strong>it</strong>h the claim that Iw,n<br />

is continuous. Nevertheless, Iw,n is continuous in the norm of C0 ( Ī), i.e., for any n ≥ 1,<br />

we can find γ ≡ γ(n) such that<br />

(6.8.2) Iw,nf C 0 (Ī) ≤ γ f C 0 (Ī), ∀f ∈ C 0 ( Ī).<br />

In fact, for any f ∈ C0 ( Ī), (3.9.5) yields<br />

(6.8.3) Iw,nf C 0 (Ī) ≤ γ |||Iw,nf|||w,n = γ max<br />

1≤j≤n |f(ξ(n)<br />

j )| ≤ γ f C 0 (Ī).<br />

This is why the natural domain for Iw,n is the space of continuous functions. Unfor-<br />

tunately, the constant γ grows w<strong>it</strong>h n.<br />

* * * * * * * * * * * *<br />

W<strong>it</strong>h the conclusion of this chapter, we have finished the first part of the book. Having<br />

established the structure of the approximating functions we intend to use, as well as their<br />

properties, we now turn our efforts toward the discretization of derivative operators, and<br />

their associated problems.


7<br />

DERIVATIVE MATRICES<br />

The derivative operator is a linear transformation from the space of polynomials into<br />

<strong>it</strong>self. In add<strong>it</strong>ion, the exact derivative of a given polynomial can be determined in<br />

a fin<strong>it</strong>e number of operations. Depending on the basis in which the polynomial is<br />

expressed, we will construct the appropriate matrices that allow the computation of <strong>it</strong>s<br />

derivative.<br />

7.1 Derivative matrices in the frequency space<br />

Let p be a polynomial in Pn, where n ≥ 1 is given. From (2.3.1), and u ′ 0 ≡ 0, we can<br />

wr<strong>it</strong>e<br />

(7.1.1) p ′ =<br />

n<br />

k=1<br />

cku ′ k =<br />

n<br />

k=0<br />

c (1)<br />

k uk.<br />

In (7.1.1), the constants c (1)<br />

k , 0 ≤ k ≤ n, are the Fourier coefficients of p′ . We would like<br />

to establish a relation between these new coefficients and the ck’s. Let us begin w<strong>it</strong>h the<br />

Jacobi case (uk = P (α,β)<br />

k , k ∈ N). For simplic<strong>it</strong>y, we just discuss the ultraspherical case<br />

(ν := α = β). Thus, we start by considering the following relation (see szegö(1939),<br />

p.84):<br />

(7.1.2) un = d<br />

dx<br />

<br />

(n + 2ν + 1) un+1<br />

(n + ν + 1)(2n + 2ν + 1) −<br />

<br />

(n + ν) un−1<br />

, n ≥ 1,<br />

(n + 2ν)(2n + 2ν + 1)


126 Polynomial Approximation of Differential Equations<br />

w<strong>it</strong>h un = P (ν,ν)<br />

n , ν > −1 and ν = −1 2 .<br />

After tedious calculation, we recover from (7.1.2) the expression of the derivatives of<br />

the ultraspherical polynomials, i.e.<br />

(7.1.3)<br />

d<br />

dx un =<br />

×<br />

[(n−1)/2] <br />

m=0<br />

<br />

Γ(n + ν + 1)<br />

Γ(n + 2ν + 1)<br />

2n − 4m + 2ν − 1<br />

n − 2m + 2ν<br />

Γ(n − 2m + 2ν + 1)<br />

Γ(n − 2m + ν)<br />

un−2m−1<br />

for any n ≥ 1 and ν > −1 w<strong>it</strong>h ν = −1 2 . The symbol [•] denotes the integer part of •.<br />

Particularly interesting is the Legendre case (ν = 0), where (7.1.3) takes the form<br />

(7.1.4)<br />

d<br />

dx Pn =<br />

[(n−1)/2] <br />

m=0<br />

(2n − 4m − 1)Pn−2m−1, n ≥ 1.<br />

Using the same arguments, recalling (1.5.10), we get in the Chebyshev case<br />

(7.1.5)<br />

d<br />

dx Tn =<br />

⎧<br />

⎪⎨<br />

T0<br />

n/2−1 <br />

2n<br />

m=0<br />

⎪⎩ nT0 + 2n<br />

T2m+1<br />

(n−1)/2<br />

<br />

m=1<br />

T2m<br />

if n = 1,<br />

if n ≥ 2 is even,<br />

if n ≥ 3 is odd.<br />

Subst<strong>it</strong>uting formula (7.1.3) into (7.1.1), we can relate the coefficients of p ′ to those of<br />

p. First of all, we note that c (1)<br />

(7.1.6) c (1)<br />

i<br />

2i + 2ν + 1<br />

=<br />

i + 2ν + 1<br />

n = 0. Then, for ν > −1 w<strong>it</strong>h ν = −1 2<br />

Γ(i + 2ν + 2)<br />

Γ(i + ν + 1)<br />

n<br />

j=i+1<br />

i+j odd<br />

<br />

,<br />

, we have<br />

Γ(j + ν + 1)<br />

Γ(j + 2ν + 1) cj,<br />

0 ≤ i ≤ n − 1.<br />

In the Legendre case (ν = 0), the above formula yields (see also the appendix in the<br />

book of gottlieb and orszag(1977))


Derivative Matrices 127<br />

(7.1.7) c (1)<br />

i<br />

= (2i + 1)<br />

n<br />

j=i+1<br />

i+j odd<br />

cj, 0 ≤ i ≤ n − 1.<br />

For the Chebyshev basis, we still get c (1)<br />

n = 0. Besides, one has<br />

(7.1.8) c (1)<br />

i =<br />

⎧<br />

n<br />

jcj if i = 0,<br />

j=1 ⎪⎨ j odd<br />

n<br />

2 jcj if 1 ≤ i ≤ n − 1.<br />

⎪⎩<br />

j=i+1<br />

i+j odd<br />

The linear mapping which transforms the vector {cj}0≤j≤n into the vector {c (1)<br />

i }0≤i≤n<br />

is clearly expressed by a (n + 1) × (n + 1) matrix which is upper triangular and has a<br />

vanishing diagonal . Therefore, all <strong>it</strong>s eigenvalues are zero. By applying k times (k ≥ 1)<br />

the derivative matrix to the vector {cj}0≤j≤n we obtain the vector {c (k)<br />

i }0≤i≤n, which<br />

represents the Fourier coefficients of the k-th derivative of the polynomial p. This implies<br />

that, since d n+1<br />

dx n+1 p ≡ 0 ∀p ∈ Pn, the n + 1 power of the matrix has all <strong>it</strong>s entries equal<br />

to zero.<br />

For 2 ≤ k ≤ n, an analytic formula relating the coefficients c (k)<br />

i , 0 ≤ i ≤ n, to the<br />

coefficients cj, 0 ≤ j ≤ n, is given in karageorghis and phillips(1989). Here<br />

we just present such a formula for k = 2, respectively for Legendre and Chebyshev<br />

expansions. In these two cases, c (2)<br />

n = c (2)<br />

n−1<br />

(7.1.9) c (2)<br />

i<br />

=<br />

(7.1.10) c (2)<br />

i<br />

<br />

i + 1<br />

2<br />

⎧<br />

=<br />

1<br />

2<br />

⎪⎨<br />

⎪⎩<br />

n <br />

j=i+2<br />

i+j even<br />

n<br />

j=2<br />

j even<br />

n<br />

j=i+2<br />

i+j even<br />

j 3 cj<br />

= 0, and<br />

(j(j + 1) − i(i + 1)) cj, 0 ≤ i ≤ n − 2,<br />

j(j 2 − i 2 )cj<br />

if i = 0,<br />

if 1 ≤ i ≤ n − 2.


128 Polynomial Approximation of Differential Equations<br />

As proposed in section 2.3, one can study the relation between the Fourier coeffi-<br />

cients of polynomials such as xp ′ (or more involved expressions) and the coefficients of<br />

p. Many examples of this type, for Legendre and Chebyshev expansions, are discussed<br />

in the appendix of gottlieb and orszag(1977).<br />

For Laguerre polynomials, relations taken from szegö(1939), p.98, lead to<br />

(7.1.11)<br />

Therefore, one obtains<br />

(7.1.12) c (1)<br />

i<br />

d<br />

dx L(α)<br />

n−1 <br />

n = −<br />

m=0<br />

= −<br />

n<br />

j=i+1<br />

L (α)<br />

m , n ≥ 1, α > −1.<br />

cj, 0 ≤ i ≤ n − 1, and c (1)<br />

n = 0.<br />

The Herm<strong>it</strong>e case is very easy to analyze. Actually, from (1.7.8), we have<br />

(7.1.13) c (1)<br />

i = 2(i + 1) ci+1, 0 ≤ i ≤ n − 1, and c (1)<br />

n = 0.<br />

An in<strong>it</strong>ial analysis suggests that the evaluation in the frequency space of the deriva-<br />

tive of the polynomial p ∈ Pn has a computational cost proportional to n 2 , since <strong>it</strong> cor-<br />

responds to a matrix-vector multiplication. A deeper study reveals an algor<strong>it</strong>hm w<strong>it</strong>h a<br />

cost only proportional to n. In fact, one easily proves the following recursion formula.<br />

First we compute c (1)<br />

n = 0 and<br />

(7.1.14) c (1)<br />

n−1<br />

= (2n + 2ν − 1)(n + ν)<br />

(n + 2ν)<br />

Then, we proceed backwards according to the relations<br />

(7.1.15) c (1)<br />

i<br />

(2i + 2ν + 1)(i + ν + 1)<br />

=<br />

i + 2ν + 1<br />

cn, ν > −1, ν = − 1<br />

2 .<br />

<br />

i + ν + 2<br />

(2i + 2ν + 5)(i + 2ν + 2) c(1)<br />

<br />

i+2 + ci+1 ,<br />

0 ≤ i ≤ n − 2, ν > −1, ν = − 1<br />

2 .


Derivative Matrices 129<br />

Similarly, for the Chebyshev and Laguerre cases, we respectively get<br />

(7.1.16) c (1)<br />

i<br />

(7.1.17) c (1)<br />

i<br />

⎧<br />

0 if i = n,<br />

=<br />

⎪⎨ 2ncn<br />

⎪⎩<br />

c (1)<br />

i+2 + 2(i + 1)ci+1 if<br />

if i = n − 1,<br />

1 ≤ i ≤ n − 2,<br />

1<br />

2 c(1)<br />

2 + c1 if i = 0,<br />

=<br />

⎧<br />

⎨ 0 if i = n,<br />

⎩<br />

0 ≤ i ≤ n − 1.<br />

c (1)<br />

i+1 − ci+1 if<br />

The Herm<strong>it</strong>e case has already a simplified form given by (7.1.13).<br />

7.2 Derivative matrices in the physical space<br />

Here we consider Pn−1, n ≥ 2, to be generated by the basis of Lagrange polynomials<br />

{l (n)<br />

j }1≤j≤n introduced in section 3.2. In view of (3.2.2), the derivative of a polynomial<br />

p ∈ Pn is obtained by evaluating<br />

(7.2.1)<br />

In particular, we have<br />

d<br />

p =<br />

dx<br />

(7.2.2) p ′ (ξ (n)<br />

i ) =<br />

n<br />

j=1<br />

n<br />

j=1<br />

p(ξ (n)<br />

j ) d<br />

dx l(n)<br />

j .<br />

d (1)<br />

ij p(ξ(n)<br />

j ), 1 ≤ i ≤ n,


130 Polynomial Approximation of Differential Equations<br />

where<br />

(7.2.3) d (1)<br />

ij :=<br />

<br />

d<br />

dx l(n)<br />

<br />

j (ξ (n)<br />

i ), 1 ≤ i ≤ n, 1 ≤ j ≤ n.<br />

Since the polynomials p and p ′ are uniquely determined by their values at the nodes<br />

ξ (n)<br />

i , 1 ≤ i ≤ n, the linear transformation (7.2.2) is equivalent to performing an exact<br />

derivative in the space Pn−1. The associated n × n matrix Dn := {d (1)<br />

ij<br />

} 1≤i≤n , allows<br />

1≤j≤n<br />

us to pass from the the vector {p(ξ (n)<br />

j )}1≤j≤n to the vector {p ′ (ξ (n)<br />

i )}1≤i≤n, w<strong>it</strong>h a<br />

computational cost proportional to n 2 .<br />

The next step is to give an explic<strong>it</strong> expression to the entries d (1)<br />

ij , 1 ≤ i ≤ n,<br />

1 ≤ j ≤ n. Computing the derivatives of the Lagrange polynomials in (3.2.4) yields<br />

(7.2.4)<br />

<br />

d<br />

dx l(n)<br />

<br />

j<br />

(x) = u′ n(x)(x − ξ (n)<br />

j ) − un(x)<br />

u ′ n(ξ (n)<br />

j ) (x − ξ (n)<br />

j ) 2<br />

After evaluation of (7.2.4) at the nodes, one obtains<br />

(7.2.5) d (1)<br />

ij =<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

u ′ n(ξ (n)<br />

i )<br />

u ′ n(ξ (n)<br />

j )<br />

u ′′ n(ξ (n)<br />

i )<br />

2 u ′ n(ξ (n)<br />

i )<br />

ξ (n)<br />

i<br />

1<br />

− ξ(n)<br />

j<br />

, x ∈ I, x = ξ (n)<br />

j , 1 ≤ j ≤ n.<br />

if i = j,<br />

if i = j.<br />

To get (7.2.5) one recalls that un(ξ (n)<br />

i ) = 0, 1 ≤ i ≤ n. The diagonal entries (i.e.,<br />

i = j) are obtained by noting that<br />

(7.2.6) lim<br />

x→ξ (n)<br />

i<br />

<br />

d<br />

dx l(n)<br />

<br />

i (x) = lim<br />

x→ξ (n)<br />

i<br />

u ′′ n(x)<br />

2 u ′ n(ξ (n)<br />

i ) = u′′ n(ξ (n)<br />

i )<br />

2 u ′ n(ξ (n)<br />

, 1 ≤ i ≤ n.<br />

i )<br />

Recalling that un is solution of a Sturm-Liouville problem, we can express the values<br />

u ′′ n(ξ (n)<br />

i ), 1 ≤ i ≤ n, in terms of u ′ (ξ (n)<br />

i ), 1 ≤ i ≤ n. This gives


Derivative Matrices 131<br />

(7.2.7) (Jacobi) d (1)<br />

ii<br />

(7.2.8) (Laguerre) d (1)<br />

ii<br />

(α + β + 2)ξ(n) i<br />

=<br />

2 (1 − (ξ (n)<br />

i ) 2 )<br />

(7.2.9) (Herm<strong>it</strong>e) d (1)<br />

ii<br />

ξ(n) i =<br />

2ξ (n)<br />

i<br />

= ξ(n)<br />

i .<br />

+ α − β<br />

, α > −1, β > −1,<br />

− α − 1<br />

, α > −1,<br />

Higher order derivative matrices are obtained by multiplying Dn by <strong>it</strong>self a su<strong>it</strong>able<br />

number of times. For example, an explic<strong>it</strong> expression of D 2 n in the Herm<strong>it</strong>e case is given<br />

in funaro and kavian(1988).<br />

In general, we denote by d (k)<br />

ij , 1 ≤ i ≤ n, 1 ≤ j ≤ n, the entries of the matrix Dk n,<br />

k ≥ 1. Of course, when k ≥ n, we have d (k)<br />

ij<br />

= 0, 1 ≤ i ≤ n, 1 ≤ j ≤ n.<br />

We can argue in the same way using the Gauss-Lobatto points. For any p ∈ Pn,<br />

n ≥ 1, we have<br />

(7.2.10) p ′ (η (n)<br />

i ) =<br />

where<br />

(7.2.11)<br />

˜ d (1)<br />

ij :=<br />

n<br />

j=0<br />

˜d (1)<br />

ij p(η(n)<br />

j ), 0 ≤ i ≤ n,<br />

<br />

d<br />

dx ˜l (n)<br />

<br />

j (η (n)<br />

i ), 0 ≤ i ≤ n, 0 ≤ j ≤ n.<br />

Similarly, we define the (n + 1) × (n + 1) matrix ˜ Dn := { ˜ d (1)<br />

ij<br />

} 0≤i≤n . Again, we denote<br />

0≤j≤n<br />

by ˜ d (k)<br />

ij , 0 ≤ i ≤ n, 0 ≤ j ≤ n, the entries of the (n + 1) × (n + 1) matrix ˜ D k n, k ≥ 1.<br />

When k ≥ n + 1, ˜ D k n is the zero matrix.<br />

The computation of the entries of ˜ Dn needs more perseverance, but <strong>it</strong> is not hard to<br />

manage. From (3.2.8), we have


132 Polynomial Approximation of Differential Equations<br />

(7.2.12) ˜ d (1)<br />

ij =<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

α − n(n + α + β + 1)<br />

2(β + 2)<br />

(−1) n n! Γ(β + 2) un(η (n)<br />

i )<br />

Γ(n + β + 1) (1 + η (n)<br />

i )<br />

(−1) n Γ(β + 2) Γ(n + α + 1)<br />

2 Γ(α + 2) Γ(n + β + 1)<br />

−(−1) n Γ(n + β + 1)<br />

n! Γ(β + 2) un(η (n)<br />

j )(1 + η (n)<br />

j )<br />

un(η (n)<br />

i )<br />

un(η (n)<br />

j )<br />

η (n)<br />

i<br />

1<br />

− η(n)<br />

j<br />

+ α − β<br />

(α + β)η (n)<br />

i<br />

2 (1 − (η (n)<br />

i ) 2 )<br />

Γ(n + α + 1)<br />

n! Γ(α + 2) un(η (n)<br />

j )(1 − η (n)<br />

j )<br />

−(−1) n Γ(α + 2) Γ(n + β + 1)<br />

2 Γ(β + 2) Γ(n + α + 1)<br />

−n! Γ(α + 2) un(η (n)<br />

i )<br />

Γ(n + α + 1) (1 − η (n)<br />

i )<br />

n(n + α + β + 1) − β<br />

2(α + 2)<br />

i = j = 0,<br />

1 ≤ i ≤ n − 1, j = 0,<br />

i = n, j = 0,<br />

i = 0, 1 ≤ j ≤ n − 1;<br />

i = j, 1 ≤ i ≤ n − 1, 1 ≤ j ≤ n − 1,<br />

1 ≤ i = j ≤ n − 1,<br />

i = n, 1 ≤ j ≤ n − 1,<br />

i = 0, j = n,<br />

1 ≤ i ≤ n − 1, j = n,<br />

i = j = n.<br />

To check (7.2.12), the reader must remember that u ′ n(η (n)<br />

j ) = 0, 1 ≤ j ≤ n − 1, while<br />

u ′ n(1) and u ′ n(−1) are respectively given by (3.1.17) and (3.1.18). Further relations,<br />

useful in this computation, are obtained from the differential equation (1.3.1). This can<br />

be used for instance to obtain the values of u ′′ n at the nodes.


Derivative Matrices 133<br />

Taking ν = 0 in (7.2.12), we recover the entries of ˜ Dn in the Legendre case:<br />

⎧<br />

(7.2.13)<br />

˜ d (1)<br />

ij =<br />

⎪⎨<br />

⎪⎩<br />

−1 4n(n + 1) i = j = 0,<br />

Pn(η (n)<br />

i )<br />

Pn(η (n)<br />

j )<br />

η (n)<br />

i<br />

1<br />

− η(n)<br />

j<br />

0 ≤ i ≤ n, 0 ≤ j ≤ n, i = j,<br />

0 1 ≤ i = j ≤ n − 1,<br />

1<br />

4n(n + 1) i = j = n.<br />

Similarly, (1.5.1) and (3.1.15) lead to the Chebyshev case (ν = −1 2 ):<br />

⎧<br />

(7.2.14)<br />

˜ d (1)<br />

ij =<br />

⎪⎨<br />

⎪⎩<br />

− 1<br />

6 (2n2 + 1) i = j = 0,<br />

1<br />

2 (−1)i /(1 + η (n)<br />

i ) 1 ≤ i ≤ n − 1, j = 0,<br />

1<br />

2 (−1)n i = n, j = 0,<br />

−2 (−1) j /(1 + η (n)<br />

j ) i = 0, 1 ≤ j ≤ n − 1,<br />

(−1) i+j<br />

η (n)<br />

i<br />

− η(n)<br />

j<br />

−η (n)<br />

i<br />

2 (1 − (η (n)<br />

i ) 2 )<br />

1 ≤ i ≤ n − 1, 1 ≤ j ≤ n − 1, i = j,<br />

1 ≤ i = j ≤ n − 1,<br />

2 (−1) j+n /(1 − η (n)<br />

j ) i = n, 1 ≤ j ≤ n − 1,<br />

− 1<br />

2 (−1)n i = 0, j = n,<br />

− 1<br />

2 (−1)i+n /(1 − η (n)<br />

i ) 1 ≤ i ≤ n − 1, j = n,<br />

1<br />

6 (2n2 + 1) i = j = n.


134 Polynomial Approximation of Differential Equations<br />

The matrices (7.2.13) and (7.2.14) have been presented for instance in gottlieb,<br />

hussaini and orszag(1984), together w<strong>it</strong>h the entries of the matrices corresponding<br />

to the second derivatives.<br />

{ ˜ d (1)<br />

ij<br />

For the Laguerre Gauss-Radau case, the entries of the n × n matrix ˜ Dn =<br />

} 0≤i≤n−1 , are obtained by (3.2.11). These are<br />

0≤j≤n−1<br />

(7.2.15) ˜ d (1)<br />

ij =<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

n − 1<br />

−<br />

α + 2<br />

n! Γ(α + 2) L (α)<br />

n (η (n)<br />

i )<br />

Γ(n + α + 1) η (n)<br />

i<br />

−Γ(n + α + 1)<br />

n! Γ(α + 2) η (n)<br />

j L(α) n (η (n)<br />

j )<br />

L (α)<br />

n (η (n)<br />

i )<br />

L (α)<br />

n (η (n)<br />

j )<br />

η (n)<br />

i<br />

2 η (n)<br />

i<br />

− α<br />

η (n)<br />

i<br />

1<br />

− η(n)<br />

j<br />

For the computation one has to recall (3.1.19) and (1.6.1).<br />

i = j = 0,<br />

1 ≤ i ≤ n − 1, j = 0,<br />

i = 0, 1 ≤ j ≤ n − 1,<br />

i = j 1 ≤ i ≤ n − 1, 1 ≤ j ≤ n − 1,<br />

1 ≤ i = j ≤ n − 1.<br />

In all the examples discussed, the algor<strong>it</strong>hm to perform the derivative in the physical<br />

space corresponds to a matrix-vector multiplication. Therefore, <strong>it</strong>s cost is proportional<br />

to n 2 . The matrices Dn and ˜ Dn, in the various cases, are full and do not display<br />

any particular property. This means that in general we cannot improve the procedure<br />

of evaluating a derivative, as suggested in the previous section. Nevertheless, for the<br />

Chebyshev case, a faster algor<strong>it</strong>hm exists. In view of the results in section 4.3, when n is<br />

a power of 2, we can go from the physical space to the frequency space (and conversely)<br />

by applying the FFT w<strong>it</strong>h a cost proportional to nlog 2n. Therefore, we can use (7.1.8),<br />

the cost of which is only proportional to n. This observation makes the Chebyshev<br />

Lagrange basis preferable, when large values of n are taken into account.


Derivative Matrices 135<br />

We conclude this section w<strong>it</strong>h a few add<strong>it</strong>ional remarks. Given the function f ∈<br />

C1 ( Ī), we can evaluate the quant<strong>it</strong>ies f(ξ(n) j ), 1 ≤ j ≤ n. After applying the matrix<br />

Dn, we obtain the vector {(Iw,nf) ′ (ξ (n)<br />

i )}1≤i≤n. For n sufficiently large, (Iw,nf) ′ is a<br />

fairly good approximation of f ′ (see (6.6.8)).<br />

Other techniques, such as fin<strong>it</strong>e-differences, can be used to compute numerically the<br />

derivative of f. Although they generate derivative matrices w<strong>it</strong>h qu<strong>it</strong>e good structure<br />

(for instance, they are banded w<strong>it</strong>h a narrow bandwidth), the results are in general very<br />

poor when compared w<strong>it</strong>h those obtained by the procedure described above. The reason<br />

is that fin<strong>it</strong>e-differences are local methods. To compute the derivative at a given point,<br />

these methods use information pertaining to a small neighborhood of the point. In<br />

contrast, spectral methods are global methods. All n nodes in the domain I are applied<br />

in the computation. For example, when f is a polynomial in Pn−1, the values at the<br />

nodes are representative of the whole function in the domain Ī, and the exact derivative<br />

is obtained. The particular distribution of the nodes optimizes the process for a general<br />

f, and the accuracy increases rapidly w<strong>it</strong>h the regular<strong>it</strong>y of the function. The reader<br />

may find other persuasive arguments supporting these considerations in boyd(1989),<br />

p.11.<br />

7.3 Boundary cond<strong>it</strong>ions in the frequency space<br />

Clearly, the procedure of evaluating the derivative q := p ′ ∈ Pn−1 of a polynomial<br />

p ∈ Pn can be inverted except to w<strong>it</strong>hin an add<strong>it</strong>ive constant. Actually, the derivative<br />

formulas of section 7.1 do not take into consideration the coefficient c0. We need an<br />

extra cond<strong>it</strong>ion. The most popular approach is to assume that p satisfies p(ξ) = σ ∈ R,<br />

where ξ ∈ Ī. This is a boundary cond<strong>it</strong>ion when ξ ∈ ∂I, i.e., ξ is an element of the<br />

boundary of Ī. In this case, we can recover the coefficients ci, 1 ≤ i ≤ n, in terms<br />

of the coefficients dj, 0 ≤ j ≤ n − 1, of the expansion q = n−1<br />

j=0 djuj. Finally, c0 is


136 Polynomial Approximation of Differential Equations<br />

determined according to the following relation:<br />

(7.3.1) c0 = σ −<br />

n<br />

k=1<br />

ck uk(ξ).<br />

This comes from evaluating (2.3.1) at the point x = ξ. Many other cond<strong>it</strong>ions are<br />

possible. For instance, one can impose <br />

pdx = σ, σ ∈ R. In the Chebyshev case, by<br />

(2.6.4), c0 is then obtained from<br />

(7.3.2) c0 = σ<br />

2 −<br />

I<br />

n<br />

k=2<br />

k even<br />

ck<br />

.<br />

1 − k2 Let us analyze a more general s<strong>it</strong>uation. We just consider the ultraspherical case.<br />

Let A and σ be real constants and ξ ∈ Ī. Given the polynomial q ∈ Pn−1, we are<br />

concerned w<strong>it</strong>h finding p ∈ Pn, such that<br />

(7.3.3)<br />

⎧<br />

⎨<br />

p ′ + Πw,n−1(Ap) = q,<br />

⎩<br />

p(ξ) = σ.<br />

The projection operator Πw,n, n ∈ N, has been introduced in section 2.4.<br />

For example, when n = 4, problem (7.3.3) is equivalent to the following system:<br />

(7.3.4)<br />

⎡<br />

A ∗ 0 ∗<br />

⎤⎡<br />

0<br />

⎢ 0<br />

⎢ 0<br />

⎣<br />

0<br />

A<br />

0<br />

0<br />

∗<br />

A<br />

0<br />

0<br />

∗<br />

A<br />

∗⎥<br />

⎢<br />

⎥⎢<br />

0⎥<br />

⎢<br />

⎦⎣<br />

∗<br />

∗ ∗ ∗ ∗ ∗<br />

c0<br />

c1<br />

c2<br />

c3<br />

c4<br />

⎤<br />

⎥<br />

⎦ =<br />

where the dk’s are the Fourier coefficients of q, and the symbol ∗ denotes the non zero<br />

entries. Due to the special structure of the matrix in (7.3.4), the polynomial p in (7.3.3)<br />

can be computed by Gauss elimination w<strong>it</strong>h a cost proportional to n 2 . However, the<br />

s<strong>it</strong>uation gets far more complicated when A : I → R is a non constant function. In<br />

this case the matrix associated to the system is full (see section 2.4).<br />

⎡<br />

⎢<br />

⎣<br />

d0<br />

d1<br />

d2<br />

d3<br />

σ<br />

⎤<br />

⎥<br />

⎦ ,


Derivative Matrices 137<br />

Next, consider second-order derivatives. Now, two add<strong>it</strong>ional cond<strong>it</strong>ions are re-<br />

quired to select a unique prim<strong>it</strong>ive function. Since there are several possibil<strong>it</strong>ies, we<br />

only consider one example. Let A, B, σ1, σ2, be real constants and q ∈ Pn−2, n ≥ 2.<br />

In the ultraspherical case, one is concerned w<strong>it</strong>h finding p ∈ Pn such that<br />

⎧<br />

(7.3.5)<br />

−p ′′ + Πw,n−2(Ap ′ + Bp) = q,<br />

⎪⎨<br />

p(−1) = σ1,<br />

⎪⎩<br />

p(1) = σ2.<br />

For instance, when n = 5, problem (7.3.5) leads to a linear system of the form<br />

(7.3.6)<br />

⎡<br />

B ∗ ∗ ∗ ∗<br />

⎤⎡<br />

∗<br />

⎢ 0<br />

⎢ 0<br />

⎢ 0<br />

⎣<br />

∗<br />

B<br />

0<br />

0<br />

∗<br />

∗<br />

B<br />

0<br />

∗<br />

∗<br />

∗<br />

B<br />

∗<br />

∗<br />

∗<br />

∗<br />

∗<br />

∗ ⎥⎢<br />

⎥⎢<br />

∗ ⎥⎢<br />

⎥⎢<br />

∗ ⎥⎢<br />

⎦⎣<br />

∗<br />

∗ ∗ ∗ ∗ ∗ ∗<br />

where the dk’s are the Fourier coefficients of q. Again, this can be solved w<strong>it</strong>h a direct<br />

procedure w<strong>it</strong>h a cost proportional to n 2 .<br />

From (1.3.2) and (1.3.3), in the ultraspherical case we have uk(1) = (−1) k uk(−1),<br />

k ∈ N. This implies<br />

(7.3.7) 2<br />

n<br />

k=0<br />

k even<br />

ck uk(1) = σ1 + σ2, 2<br />

c0<br />

c1<br />

c2<br />

c3<br />

c4<br />

c5<br />

⎤<br />

⎥<br />

⎦<br />

n<br />

=<br />

k=1<br />

k odd<br />

⎡<br />

⎢<br />

⎣<br />

d0<br />

d1<br />

d2<br />

d3<br />

σ1<br />

σ2<br />

⎤<br />

⎥<br />

⎥,<br />

⎥<br />

⎦<br />

ck uk(1) = σ1 − σ2.<br />

Therefore, when A = 0, (7.3.6) can be decoupled into the subsystems<br />

(7.3.8)<br />

⎡<br />

B<br />

⎣ 0<br />

∗<br />

B<br />

⎤⎡<br />

∗<br />

∗⎦<br />

⎣<br />

∗ ∗ ∗<br />

c0<br />

⎤ ⎡<br />

c2 ⎦ = ⎣ d0<br />

d2<br />

⎤<br />

⎦,<br />

⎡<br />

B<br />

⎣ 0<br />

∗<br />

B<br />

⎤⎡<br />

∗<br />

∗ ⎦⎣<br />

∗ ∗ ∗<br />

c1<br />

⎤ ⎡<br />

c3 ⎦ =<br />

c4<br />

σ1 + σ2<br />

c5<br />

⎣ d1<br />

d3<br />

σ1 − σ2<br />

This is because c (2)<br />

i , 0 ≤ i ≤ n − 2, is a linear combination of cj, i + 2 ≤ j ≤ n, where<br />

i+j is even (see for instance (7.1.9) and (7.1.10)). Following this idea, a fast algor<strong>it</strong>hm<br />

for the solution of (7.3.5) is presented in clenshaw(1957), for the Chebyshev case.<br />

Further hints are given in canuto, hussaini, quarteroni and zang(1988), p.129,<br />

and boyd(1989), p.380.<br />

⎤<br />

⎦.


138 Polynomial Approximation of Differential Equations<br />

Different boundary cond<strong>it</strong>ions can be examined. For B > 0, typical cond<strong>it</strong>ions are<br />

obtained by imposing p ′ (−1) = σ1, p ′ (1) = σ2. The last two rows in (7.3.6) have to be<br />

changed accordingly.<br />

One can also assume that A and B are non constant functions in I. This implies<br />

that the matrix relative to the system (7.3.5) is full. Moreover, the expression of the<br />

Fourier coefficients of the term Πw,n−2(Ap ′ +Bp), n ≥ 2, is qu<strong>it</strong>e involved (see section<br />

2.4). As we see in the next section, <strong>it</strong> is more natural to treat these kind of problems<br />

in the physical space. Starting from a different approach, further methods to impose<br />

boundary cond<strong>it</strong>ions in the frequency space will be analyzed in section 9.4. Suggestions<br />

for the treatment of the Laguerre and Herm<strong>it</strong>e cases are given in section 9.5.<br />

7.4 Boundary cond<strong>it</strong>ions in the physical space<br />

It is standard to work w<strong>it</strong>h Gauss-Lobatto nodes in the Jacobi case and Gauss-Radau<br />

nodes in the Laguerre case, since these actually include the boundary points of the<br />

domain<br />

Ī. Applications of the Gauss-Radau nodes in (3.1.13) and (3.1.14) have been<br />

considered in canuto and quarteroni(1982b) for boundary cond<strong>it</strong>ions on one side of<br />

the interval [−1,1].<br />

We give a survey of the cases that can occur in the discretization of linear boundary-<br />

value problems in one dimension. The study of the generating equations and of the<br />

behavior of the approximated solutions, when the number of nodes increases, are the<br />

subject of chapter nine.<br />

As usual, we first consider the Jacobi case. The problem of inverting the derivative<br />

operator in the space of polynomials can be formulated as follows. Let us assume that<br />

ξ = −1 (similar arguments apply when ξ = 1). Let q be a polynomial in Pn−1, n ≥ 1.<br />

We are concerned w<strong>it</strong>h finding p ∈ Pn such that p ′ = q and p(ξ) = σ, where σ ∈ R.<br />

Evaluating these expressions at the nodes, we end up w<strong>it</strong>h the following set of equations:


Derivative Matrices 139<br />

(7.4.1)<br />

⎧<br />

⎨<br />

⎩<br />

p ′ (η (n)<br />

i ) = q(η (n)<br />

i ) 1 ≤ i ≤ n,<br />

p(η (n)<br />

0 ) = σ.<br />

This is equivalent to a linear system in the unknowns p(η (n)<br />

i ), 0 ≤ i ≤ n. For example,<br />

when n = 3, w<strong>it</strong>h the notations of section 7.2, we get<br />

(7.4.2)<br />

⎡<br />

⎢<br />

⎣<br />

1 0 0 0<br />

˜d (1)<br />

10<br />

˜d (1)<br />

20<br />

˜d (1)<br />

30<br />

˜d (1)<br />

11<br />

˜d (1)<br />

21<br />

˜d (1)<br />

31<br />

˜d (1)<br />

12<br />

˜d (1)<br />

22<br />

˜d (1)<br />

32<br />

˜d (1)<br />

13<br />

˜d (1)<br />

23<br />

˜d (1)<br />

33<br />

⎤⎡<br />

p(η<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎦⎢<br />

⎣<br />

(n)<br />

0 )<br />

p(η (n)<br />

1 )<br />

p(η (n)<br />

⎤<br />

⎥<br />

2 ) ⎥<br />

⎦<br />

p(η (n)<br />

3 )<br />

=<br />

⎡<br />

σ<br />

⎢<br />

⎢q(η<br />

⎢<br />

⎣<br />

(n)<br />

1 )<br />

q(η (n)<br />

⎤<br />

⎥<br />

⎥.<br />

⎥<br />

2 ) ⎥<br />

⎦<br />

q(η (n)<br />

3 )<br />

If r ∈ Pn is the solution of (7.4.2) w<strong>it</strong>h σ = 0, we note that r satisfies<br />

(7.4.3)<br />

⎡<br />

⎢<br />

⎣<br />

˜d (1)<br />

11<br />

˜d (1)<br />

21<br />

˜d (1)<br />

31<br />

˜d (1)<br />

12<br />

˜d (1)<br />

22<br />

˜d (1)<br />

32<br />

˜d (1)<br />

13<br />

˜d (1)<br />

23<br />

˜d (1)<br />

33<br />

⎤⎡<br />

r(η<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎦⎣<br />

(n)<br />

1 )<br />

r(η (n)<br />

⎤<br />

⎥<br />

2 ) ⎥<br />

⎦ =<br />

r(η (n)<br />

3 )<br />

⎡<br />

q(η<br />

⎢<br />

⎣<br />

(n)<br />

1 )<br />

q(η (n)<br />

⎤<br />

⎥<br />

2 ) ⎥<br />

⎦ .<br />

q(η (n)<br />

3 )<br />

Therefore, we reduced the in<strong>it</strong>ial (n + 1) × (n + 1) system to a n × n system. The<br />

final solution is then obtained by setting p = r + σ.<br />

Another interesting problem is stated as follows. For q ∈ Pn and γ ∈ R, γ = 0,<br />

find p ∈ Pn such that<br />

(7.4.4)<br />

⎧<br />

⎨<br />

⎩<br />

p ′ (η (n)<br />

i ) = q(η (n)<br />

i ) 1 ≤ i ≤ n,<br />

p ′ (η (n)<br />

0 ) + γp(η (n)<br />

0 ) = q(η (n)<br />

0 ) + γσ.<br />

We observe that now the boundary constraint is not exactly imposed. Actually, we<br />

are trying to enforce both the equation and the boundary cond<strong>it</strong>ion at the point η (n)<br />

0 .<br />

Using the Lagrange polynomials introduced in theorem 3.2.1, an equivalent formulation<br />

is:


140 Polynomial Approximation of Differential Equations<br />

(7.4.5) p ′ + γ[p(η (n)<br />

0 ) − σ] ˜l (n)<br />

0<br />

= q.<br />

This kind of approach was proposed in funaro and gottlieb(1988) for Chebyshev<br />

nodes. In terms of matrices, this is equivalent to (we take for instance n = 3)<br />

(7.4.6)<br />

⎡<br />

⎢<br />

⎣<br />

˜d (1)<br />

00 + γ d ˜(1) 01<br />

˜d (1)<br />

10<br />

˜d (1)<br />

20<br />

˜d (1)<br />

30<br />

˜d (1)<br />

11<br />

˜d (1)<br />

21<br />

˜d (1)<br />

31<br />

˜d (1)<br />

02<br />

˜d (1)<br />

12<br />

˜d (1)<br />

22<br />

˜d (1)<br />

32<br />

˜d (1)<br />

03<br />

˜d (1)<br />

13<br />

˜d (1)<br />

23<br />

˜d (1)<br />

33<br />

⎤⎡<br />

p(η<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎦⎣<br />

(n)<br />

0 )<br />

p(η (n)<br />

1 )<br />

p(η (n)<br />

⎤<br />

⎥<br />

2 ) ⎥<br />

⎦<br />

p(η (n)<br />

3 )<br />

=<br />

⎡<br />

q(η<br />

⎢<br />

⎣<br />

(n)<br />

0 ) + γσ<br />

q(η (n)<br />

1 )<br />

q(η (n)<br />

⎤<br />

⎥<br />

⎥.<br />

2 ) ⎥<br />

⎦<br />

q(η (n)<br />

3 )<br />

If γ = 0, (7.4.4) is not a well-posed problem, because the determinant of the corre-<br />

sponding matrix vanishes. Also note that, if q ∈ Pn−1, problems (7.4.1) and (7.4.4) are<br />

equivalent. Indeed, in this case, the n cond<strong>it</strong>ions p ′ (η (n)<br />

i ) = q(η (n)<br />

i ), 1 ≤ i ≤ n, imply<br />

that p ′ ≡ q in Ī, hence p′ (η (n)<br />

0 ) = q(η (n)<br />

0 ). Since γ = 0, we obtain the equivalence.<br />

Let A : Ī → R be a continuous function. A generalization of problem (7.4.1) is<br />

obtained by setting<br />

(7.4.7)<br />

⎧<br />

⎨<br />

⎩<br />

p ′ (η (n)<br />

i ) + (Ap)(η (n)<br />

i ) = q(η (n)<br />

i ) 1 ≤ i ≤ n,<br />

p(η (n)<br />

0 ) = σ,<br />

where q ∈ Pn−1 and σ ∈ R.<br />

The corresponding system is obtained by adding to the matrix of system (7.4.1), the<br />

(n + 1) × (n + 1) diagonal matrix diag{0, A(η (n)<br />

1 ), · · · , A(η (n)<br />

n )}.<br />

Problem (7.4.4) is generalized in a similar way. In particular, (7.4.5) becomes<br />

(7.4.8) p ′ + Iw,n(Ap) + γ[p(η (n)<br />

0 ) − σ] ˜l (n)<br />

0<br />

= q,<br />

where q ∈ Pn, σ ∈ R, and γ ∈ R, γ = 0. Now, we must add to the matrix relative to<br />

(7.4.4), the (n + 1) × (n + 1) diagonal matrix diag{A(η (n)<br />

0 ), A(η (n)<br />

1 ), · · · , A(η (n)<br />

n )}.


Derivative Matrices 141<br />

Let us consider now second-order problems. Let q be a polynomial in Pn−2, n ≥ 2.<br />

For σ1, σ2 ∈ R, one is concerned w<strong>it</strong>h finding p ∈ Pn such that<br />

(7.4.9)<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

−p ′′ (η (n)<br />

i ) = q(η (n)<br />

i ) 1 ≤ i ≤ n − 1,<br />

p(η (n)<br />

0 ) = σ1,<br />

p(η (n)<br />

n ) = σ2.<br />

This corresponds to find the solution of a linear system. We show the case n = 3:<br />

(7.4.10)<br />

⎡<br />

⎢<br />

⎣<br />

1 0 0 0<br />

− ˜ d (2)<br />

10 − ˜ d (2)<br />

11 − ˜ d (2)<br />

12 − ˜ d (2)<br />

13<br />

− ˜ d (2)<br />

20 − ˜ d (2)<br />

21 − ˜ d (2)<br />

22 − ˜ d (2)<br />

23<br />

0 0 0 1<br />

⎤⎡<br />

p(η<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎦⎢<br />

⎣<br />

(n)<br />

0 )<br />

p(η (n)<br />

1 )<br />

p(η (n)<br />

⎤<br />

⎥<br />

2 ) ⎥<br />

⎦<br />

p(η (n)<br />

3 )<br />

=<br />

⎡<br />

σ1<br />

⎢<br />

⎢q(η<br />

⎢<br />

⎣<br />

(n)<br />

1 )<br />

q(η (n)<br />

⎥<br />

⎥.<br />

⎥<br />

2 ) ⎥<br />

⎦<br />

Let r be the solution of (7.4.10) w<strong>it</strong>h σ1 = σ2 = 0. Thus, we can eliminate two<br />

unknowns and reduce the system to<br />

(7.4.11)<br />

⎡<br />

⎣ − ˜ d (2)<br />

11 − ˜ d (2) ⎤⎡<br />

12<br />

⎦<br />

− ˜ d (2)<br />

21 − ˜ d (2)<br />

22<br />

⎣ r(η(n)<br />

⎤<br />

1 )<br />

⎦ =<br />

r(η (n)<br />

2 )<br />

⎡<br />

⎣ q(η(n)<br />

⎤<br />

1 )<br />

⎦.<br />

q(η (n)<br />

2 )<br />

Finally, we recover p from the relation p(x) = r(x) + 1<br />

2 (1 − x)σ1 + 1<br />

2 (1 + x)σ2, ∀x ∈ Ī.<br />

Of course, this procedure applies for any n ≥ 2. We note that the matrix in (7.4.11) is<br />

not the square of some first derivative matrix.<br />

Problem (7.4.9) is generalized as follows:<br />

(7.4.12)<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

−p ′′ (η (n)<br />

i ) + (Bp ′ + Ap)(η (n)<br />

i ) = q(η (n)<br />

i ) 1 ≤ i ≤ n − 1,<br />

p(η (n)<br />

0 ) = σ1,<br />

p(η (n)<br />

n ) = σ2,<br />

where A and B are continuous functions in Ī.<br />

σ2<br />


142 Polynomial Approximation of Differential Equations<br />

Different boundary cond<strong>it</strong>ions are possible. A classical problem is to find p ∈ Pn<br />

such that<br />

(7.4.13)<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

−p ′′ (η (n)<br />

i ) + µp(η (n)<br />

i ) = q(η (n)<br />

i ) 1 ≤ i ≤ n − 1,<br />

−p ′ (η (n)<br />

0 ) = −σ1,<br />

p ′ (η (n)<br />

n ) = σ2,<br />

where q ∈ Pn−2, σ1,σ2 ∈ R, and µ > 0.<br />

We changed the sign at relation p ′ (η (n)<br />

0 ) = σ1 for a reason that will be understood in<br />

section 8.2. The matrix associated to (7.4.13), for n = 3, takes the form<br />

(7.4.14)<br />

⎡<br />

⎢<br />

⎣<br />

− ˜ d (1)<br />

00<br />

− ˜ d (1)<br />

01<br />

− ˜ d (1)<br />

02<br />

− ˜ d (2)<br />

10 − ˜ d (2)<br />

11 + µ − ˜ d (2)<br />

12<br />

− ˜ d (2)<br />

20<br />

˜d (1)<br />

30<br />

− ˜ d (2)<br />

21<br />

˜d (1)<br />

31<br />

− ˜ d (1)<br />

03<br />

− ˜ d (2)<br />

13<br />

− ˜ d (2)<br />

22 + µ − ˜ d (2)<br />

23<br />

˜d (1)<br />

32<br />

˜d (1)<br />

33<br />

⎤⎡<br />

p(η<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎦⎣<br />

(n)<br />

0 )<br />

p(η (n)<br />

1 )<br />

p(η (n)<br />

⎤<br />

⎥<br />

2 ) ⎥<br />

⎦<br />

p(η (n)<br />

3 )<br />

=<br />

⎡<br />

−σ1<br />

⎢<br />

⎢q(η<br />

⎢<br />

⎣<br />

(n)<br />

1 )<br />

q(η (n)<br />

⎤<br />

⎥<br />

⎥.<br />

⎥<br />

2 ) ⎥<br />

⎦<br />

This s<strong>it</strong>uation can be easily generalized to mixed type boundary cond<strong>it</strong>ions, such as<br />

(7.4.15)<br />

⎧<br />

⎨τ1p<br />

′ (−1) + τ2p(−1) = σ1<br />

⎩<br />

τ3p ′ (1) + τ4p(1) = σ2<br />

τi ∈ R, 1 ≤ i ≤ 4.<br />

Here τ1 and τ2 (as well as τ3 and τ4) do not vanish simultaneously.<br />

A su<strong>it</strong>able modification furnishes another formulation. Let q ∈ Pn, σ1,σ2 ∈ R, µ > 0<br />

and γ ∈ R w<strong>it</strong>h γ = 0. Then, we seek p ∈ Pn such that<br />

(7.4.16)<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

−p ′′ (η (n)<br />

i ) + µp(η (n)<br />

i ) = q(η (n)<br />

i ) 1 ≤ i ≤ n − 1,<br />

−p ′′ (η (n)<br />

0 ) + µp(η (n)<br />

0 ) − γp ′ (η (n)<br />

0 ) = q(η (n)<br />

0 ) − γσ1,<br />

−p ′′ (η (n)<br />

n ) + µp(η (n)<br />

n ) + γp ′ (η (n)<br />

n ) = q(η (n)<br />

n ) + γσ2.<br />

σ2


Derivative Matrices 143<br />

This is equivalent to<br />

(7.4.17) −p ′′ + µp − γ[p ′ (η (n)<br />

0 ) − σ1] ˜ l (n)<br />

0 + γ[p′ (η (n)<br />

n ) − σ2] ˜ l (n)<br />

n = q.<br />

This approach was introduced in funaro(1986) for Legendre nodes, and in funaro<br />

(1988) for Chebyshev nodes. For n = 3, we present the relative linear system<br />

(7.4.18)<br />

⎡<br />

⎢<br />

⎣<br />

− ˜ d (2)<br />

00 − γ ˜ d (1)<br />

00 + µ − ˜ d (2)<br />

01 − γ ˜ d (1)<br />

01<br />

− ˜ d (2)<br />

10<br />

− ˜ d (2)<br />

20<br />

− ˜ d (2)<br />

30 + γ ˜ d (1)<br />

30<br />

⎡<br />

p(η<br />

⎢<br />

× ⎢<br />

⎣<br />

(n)<br />

0 )<br />

p(η (n)<br />

1 )<br />

p(η (n)<br />

⎤<br />

⎥<br />

2 ) ⎥<br />

⎦<br />

− ˜ d (2)<br />

02 − γ ˜ d (1)<br />

02<br />

− ˜ d (2)<br />

11 + µ − ˜ d (2)<br />

12<br />

− ˜ d (2)<br />

21<br />

− ˜ d (2)<br />

31 + γ ˜ d (1)<br />

31<br />

p(η (n)<br />

3 )<br />

=<br />

⎡<br />

⎢<br />

⎣<br />

− ˜ d (2)<br />

03 − γ ˜ d (1)<br />

03<br />

− ˜ d (2)<br />

13<br />

− ˜ d (2)<br />

22 + µ − ˜ d (2)<br />

23<br />

− ˜ d (2)<br />

32 + γ ˜ d (1)<br />

32 − ˜ d (2)<br />

33 + γ ˜ d (1)<br />

33<br />

q(η (n)<br />

0 ) − γσ1<br />

q(η (n)<br />

1 )<br />

q(η (n)<br />

2 )<br />

q(η (n)<br />

3 ) + γσ2<br />

⎤<br />

⎥<br />

⎥.<br />

⎥<br />

⎦<br />

⎤<br />

⎥<br />

+ µ ⎦<br />

The reason for using these type of boundary cond<strong>it</strong>ions will be explained in section 9.4.<br />

Alternative techniques are presented in canuto(1986).<br />

We now investigate fourth-order problems. A classical example is to find p ∈ Pn+2,<br />

n ≥ 2, such that<br />

(7.4.19)<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

p IV (η (n)<br />

i ) = q(η (n)<br />

i ) 1 ≤ i ≤ n − 1,<br />

p(η (n)<br />

0 ) = σ1, p(η (n)<br />

n ) = σ2,<br />

p ′ (η (n)<br />

0 ) = σ3, p ′ (η (n)<br />

n ) = σ4,<br />

where q ∈ Pn−2 and σk ∈ R, 1 ≤ k ≤ 4. For the purpose of wr<strong>it</strong>ing the associated<br />

linear system, we construct the unique polynomial s ∈ P3 that satisfies the same<br />

boundary cond<strong>it</strong>ions as p:


144 Polynomial Approximation of Differential Equations<br />

(7.4.20) s(x) := 1<br />

<br />

2σ1 + 2σ2 + σ3 − σ4 + (3σ2 − 3σ1 − σ3 − σ4)x<br />

4<br />

+ (σ4 − σ3)x 2 + (σ1 − σ2 + σ3 + σ4)x 3<br />

, ∀x ∈ Ī.<br />

Then, r := p − s ∈ Pn+2, is expressed as follows (see also funaro and heinrichs<br />

(1990)):<br />

(7.4.21) r(x) =<br />

n−1 <br />

j=1<br />

[p(η (n)<br />

j ) − s(η (n)<br />

j )]<br />

1 − x 2<br />

1 − (η (n)<br />

j ) 2<br />

˜ l (n)<br />

j (x), ∀x ∈ Ī,<br />

where we noted that r(±1) = r ′ (±1) = 0. Deriving this relation four times and<br />

evaluating at the nodes, finally leads to the (n − 1) × (n − 1) system<br />

(7.4.22)<br />

n−1 <br />

j=1<br />

1<br />

1 − (η (n)<br />

j ) 2<br />

<br />

[1 − (η (n)<br />

i ) 2 ] ˜ d (4)<br />

ij − 8η(n) ˜d i<br />

(3)<br />

ij − 12 ˜ d (2)<br />

<br />

ij r(η (n)<br />

j ) = q(η (n)<br />

i ),<br />

Once r is computed, we determine p from the expression (7.4.20).<br />

Let us analyze another example. We wish to find p ∈ Pn+2, n ≥ 2, such that<br />

(7.4.23)<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

p IV (η (n)<br />

i ) = q(η (n)<br />

i ) 1 ≤ i ≤ n − 1,<br />

p(η (n)<br />

0 ) = σ1, p(η (n)<br />

n ) = σ2,<br />

p ′′ (η (n)<br />

0 ) = σ3, p ′′ (η (n)<br />

n ) = σ4,<br />

1 ≤ i ≤ n − 1.<br />

where q ∈ Pn−2 and σk ∈ R, 1 ≤ k ≤ 4. Again, <strong>it</strong> is sufficient to solve a (n−1)×(n−1)<br />

linear system in the unknowns r(η (n)<br />

j ), 1 ≤ j ≤ n − 1, where r := p − s and s ∈ P3<br />

is now given by<br />

(7.4.24) s(x) := 1<br />

<br />

6σ1 + 6σ2 − 3σ3 − 3σ4 + (6σ2 − 6σ1 + σ3 − σ4)x<br />

12<br />

+ (3σ3 + 3σ4)x 2 + (σ4 − σ3)x 3<br />

, ∀x ∈ Ī.


Derivative Matrices 145<br />

The matrix is given by squaring the equivalent of the matrix in (7.4.11), for a general<br />

n ≥ 2. In fact, (7.4.23) can be decoupled, w<strong>it</strong>h the subst<strong>it</strong>ution ˆp := p ′′ , into two<br />

consecutive second order systems.<br />

We finally conclude w<strong>it</strong>h some examples relative to unbounded domains. Let us<br />

consider first the case I =]0,+∞[. Here, problems are formulated in the space of<br />

Laguerre functions (see sections 6.7 and 9.5). For Q ∈ Sn−2, n ≥ 2, σ ∈ R and<br />

µ > 0, we are concerned w<strong>it</strong>h finding the solution P ∈ Sn−1 of the problem<br />

(7.4.25)<br />

⎧<br />

⎨<br />

⎩<br />

−P ′′ (η (n)<br />

i ) + µP(η (n)<br />

i ) = Q(η (n)<br />

i ), 1 ≤ i ≤ n − 1,<br />

P(η (n)<br />

0 ) = σ,<br />

where the η (n)<br />

i ’s are the Laguerre Gauss-Radau nodes (see section 3.6). By setting<br />

q(x) := Q(x)e x , p(x) := P(x)e x , ∀x ∈ [0,+∞[, we get an equivalent formulation in<br />

the space of polynomials<br />

(7.4.26)<br />

⎧<br />

⎨<br />

⎩<br />

−p ′′ (η (n)<br />

i ) + 2p ′ (η (n)<br />

i ) + (µ − 1)p(η (n)<br />

i ) = q(η (n)<br />

i ), 1 ≤ i ≤ n − 1,<br />

p(η (n)<br />

0 ) = σ.<br />

This leads to a n×n linear system. The entries of the corresponding matrix are obtained<br />

in the usual way. Similar considerations hold for other types of boundary cond<strong>it</strong>ions. We<br />

remark that we only have one boundary point for a second-order equation. Nevertheless,<br />

problem (7.4.25) adm<strong>it</strong>s a unique solution. Actually, another cond<strong>it</strong>ion is implic<strong>it</strong>ly<br />

assumed by observing that P decays to zero at infin<strong>it</strong>y exponentially fast. More<br />

details are given in section 9.5.<br />

For the Herm<strong>it</strong>e case there are no boundary points. Let us discuss an example.<br />

The generating problem will be given in section 10.2. Let Q ∈ Sn−1, n ≥ 1, be a<br />

Herm<strong>it</strong>e function and µ > 0. We want to find P ∈ Sn−1 such that<br />

(7.4.27) −P ′′ (ξ (n)<br />

i ) − 2ξ (n)<br />

i P ′ (ξ (n)<br />

i ) + µP(ξ (n)<br />

i ) = Q(ξ (n)<br />

i ), 1 ≤ i ≤ n,


146 Polynomial Approximation of Differential Equations<br />

where the ξ (n)<br />

i ’s are the zeroes of Hn. Boundary cond<strong>it</strong>ions are replaced by the relation<br />

limx→±∞ P(x) = 0. By the subst<strong>it</strong>ution of q(x) := Q(x)ex2, p(x) := P(x)ex2 , x ∈ R,<br />

into (7.4.27) we recover the corresponding problem in Pn−1, i.e.<br />

(7.4.28) −p ′′ (ξ (n)<br />

i ) + 2ξ (n)<br />

i p′ (ξ (n)<br />

i ) + µ p(ξ (n)<br />

i ) = q(ξ (n)<br />

i ), 1 ≤ i ≤ n.<br />

This leads to a non singular n × n linear system in the unknowns p(ξ (n)<br />

j ), 1 ≤ j ≤ n.<br />

7.5 Derivatives of scaled functions<br />

Following the suggestions given in funaro(1990a) and in section 1.6, w<strong>it</strong>h the help of<br />

the scaling function (1.6.13), we can in part reduce the problems associated w<strong>it</strong>h the<br />

numerical evaluation of high degree Laguerre or Herm<strong>it</strong>e polynomials.<br />

For example, starting from the matrix in (7.2.15), we define a new matrix by setting<br />

(7.5.1)<br />

d ˆ(1) ij := ˜ d (1)<br />

ij S(α) n (η (n)<br />

i )<br />

<br />

S (α)<br />

n (η (n)<br />

−1 j ) , 0 ≤ i ≤ n − 1, 0 ≤ j ≤ n − 1.<br />

Let p be a polynomial in Pn−1, n ≥ 1. Then we define ˆp := S (α)<br />

n p. Now, one has<br />

(7.5.2)<br />

n−1 <br />

j=0<br />

Thus, the matrix ˆ Dn := { ˆ d (1)<br />

ij } 0≤i≤n−1<br />

0≤j≤n−1<br />

ˆd (1)<br />

ij ˆp(η(n) j ) = S (α)<br />

n (η (n)<br />

i ) p ′ (η (n)<br />

i ), 0 ≤ i ≤ n − 1.<br />

acts on the scaled polynomial ˆp and leads to<br />

<strong>it</strong>s scaled derivative. The values at the nodes of these functions are in general more<br />

appropriate for numerical purposes. Moreover, let us better examine the entries of ˆ Dn.<br />

From (1.6.12), (1.6.13) and (7.5.1), we have


Derivative Matrices 147<br />

(7.5.3) ˆ d (1)<br />

ij =<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

n − 1<br />

−<br />

α + 2<br />

α + 1<br />

η (n)<br />

i<br />

i = j = 0,<br />

ˆL (α)<br />

n (η (n)<br />

i ) 1 ≤ i ≤ n − 1, j = 0,<br />

−1<br />

(α + 1) η (n)<br />

j<br />

ˆL (α)<br />

n (η (n)<br />

i )<br />

ˆL (α)<br />

n (η (n)<br />

j )<br />

− α<br />

η (n)<br />

i<br />

2 η (n)<br />

i<br />

ˆL (α)<br />

n (η (n)<br />

j )<br />

η (n)<br />

i<br />

1<br />

− η(n)<br />

j<br />

i = 0, 1 ≤ j ≤ n − 1,<br />

1 ≤ i ≤ n − 1, 1 ≤ j ≤ n − 1, i = j,<br />

1 ≤ i = j ≤ n − 1.<br />

All the coefficients are now computed in terms of the scaled Laguerre functions. Second-<br />

order derivative operators are obtained by squaring ˆ Dn. The corresponding entries are<br />

denoted by ˆ d (2)<br />

ij , 0 ≤ i ≤ n − 1, 0 ≤ j ≤ n − 1.<br />

It is clear that problem (7.4.26) is equivalent to the following one:<br />

(7.5.4)<br />

⎧<br />

n−1<br />

⎪⎨<br />

<br />

− ˆ d (2)<br />

ij + 2 ˆ d (1)<br />

<br />

ij + (µ − 1)δij ˆp(η (n)<br />

j ) = ˆq(η (n)<br />

i ), 1 ≤ i ≤ n − 1,<br />

⎪⎩<br />

j=0<br />

where ˆq := S (α)<br />

n q.<br />

ˆp(η (n)<br />

0 ) = σ S (α)<br />

n (0),<br />

Hence, we obtain a linear system in the unknowns ˆp(η (n)<br />

j ), 0 ≤ j ≤ n − 1, w<strong>it</strong>h the<br />

advantage of having now a matrix which is less affected by rounding errors.<br />

To evaluate the weighted norm of p, <strong>it</strong> is not necessary to go back to the quant<strong>it</strong>ies<br />

p(η (n)<br />

j ), 0 ≤ j ≤ n − 1. We can instead use (3.10.7).<br />

In the Herm<strong>it</strong>e case, the entries in (7.2.5) are modified as follows:


148 Polynomial Approximation of Differential Equations<br />

(7.5.5)<br />

d ˆ(1) ij := d (1)<br />

ij S(−1/2)<br />

n/2 ([ξ (n)<br />

i ] 2 )<br />

(7.5.6) d ˆ(1) ij := d (1)<br />

ij ξ(n) i S (1/2)<br />

(n−1)/2 ([ξ(n) i ] 2 )<br />

By (1.7.11) and (1.7.12), the above relations imply<br />

(7.5.7)<br />

ˆ d (1)<br />

ij =<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

ξ (n)<br />

i<br />

ξ (n)<br />

j<br />

ξ (n)<br />

i<br />

ˆH ′ n(ξ (n)<br />

i )<br />

ˆH ′ n(ξ (n)<br />

j )<br />

<br />

S (−1/2)<br />

n/2 ([ξ (n)<br />

j ] 2 −1 )<br />

<br />

ξ (n)<br />

j S(1/2)<br />

(n−1)/2 ([ξ(n) j ] 2 −1 )<br />

ξ (n)<br />

i<br />

1<br />

− ξ(n)<br />

j<br />

if i = j,<br />

if i = j.<br />

if n is even,<br />

if n is odd,<br />

1 ≤ i ≤ n, 1 ≤ j ≤ n.<br />

Therefore, we transform (7.4.28) to an equivalent problem where the unknown is the<br />

polynomial p multiplied by the scaling function.<br />

7.6 Numerical solution<br />

Various techniques are available for the numerical solution of the systems presented in<br />

the previous sections. Basic algor<strong>it</strong>hms are discussed in all the comprehensive books<br />

on numerical analysis. We mention for instance fox(1964), ralston(1965), isaac-<br />

son and keller(1966), dahlquist and björck(1974), hageman and young(1981),<br />

comincioli(1990).<br />

To take full advantage of the accurate results obtainable w<strong>it</strong>h spectral methods,<br />

we always suggest that one operates w<strong>it</strong>h programs wr<strong>it</strong>ten in double precision (i.e., 64<br />

b<strong>it</strong>s for a real number storage).<br />

In this book, we are only concerned w<strong>it</strong>h the discretization of differential equations<br />

in one variable. For this reason, the dimension n of the matrices is in general small.


Derivative Matrices 149<br />

Only in a very few large-scale computations n is bigger than 100 (over 200 in extreme<br />

cases). In most of the applications n does not exceed 30. In this case <strong>it</strong> is convenient<br />

to use a direct method, such as Gauss elimination w<strong>it</strong>h maximal pivots, rather than<br />

an <strong>it</strong>erative method. Direct factorization (such as LU decompos<strong>it</strong>ion) is recommended<br />

when the same system has to be solved several times w<strong>it</strong>h different data sets. It is well<br />

known that the global cost of these procedures is proportional to n 3 .<br />

As we shall see in chapter thirteen, for partial differential equations, the derivative<br />

matrices in each variable have basically the same structure as those described here. This<br />

is because spectral methods are usually applied to boundary value problems, defined<br />

on domains obtained by direct product of intervals (rectangles, cubes, and so forth). In<br />

this context, to save both computer memory and CPU time, an <strong>it</strong>erative procedure is<br />

often preferable to a direct approach. A survey of the most popular <strong>it</strong>erative algor<strong>it</strong>hms<br />

in spectral methods is given in canuto, hussaini, quarteroni and zang(1988),<br />

chapter 5, and boyd(1989), chapter 12. The development of these techniques is based<br />

on the analysis of problems in one variable. Therefore, a detailed study of the simplest<br />

s<strong>it</strong>uations will be helpful in finding new strategies.<br />

A lot of problems are defined in the physical space (see section 7.4). The major<br />

drawback, however, is that the corresponding matrices are full and non symmetric. W<strong>it</strong>h<br />

the exception of the Chebyshev case, where the FFT algor<strong>it</strong>hm considerably reduces<br />

the amount of operations (see sections 4.3 and 7.2), the cost of a matrix-vector multi-<br />

plication is proportional to n 2 . To be compet<strong>it</strong>ive, an <strong>it</strong>erative method has to reach an<br />

accurate solution in very few <strong>it</strong>erations. To achieve this goal, <strong>it</strong> is imperative to work<br />

w<strong>it</strong>h precond<strong>it</strong>ioned matrices. How to construct and implement this kind of matrix is<br />

discussed in the next chapter. For the moment, let us briefly recall some basic facts.<br />

Let D be a n×n matrix (n ≥ 1) and ¯q a given vector in R n . We want to determine<br />

the solution ¯p ∈ R n of the linear system<br />

(7.6.1) D¯p = ¯q, where det(D) = 0.<br />

A one-step <strong>it</strong>erative method to evaluate ¯p is obtained by introducing a new n × n non<br />

singular matrix R and defining an in<strong>it</strong>ial guess ¯p (0) ∈ R n . Successively, one constructs


150 Polynomial Approximation of Differential Equations<br />

the sequence of vectors such that<br />

(7.6.2) ¯p (k+1)<br />

where I is the ident<strong>it</strong>y matrix.<br />

:= (I − R −1 D)¯p (k) + R −1 ¯q, k ∈ N,<br />

Denote by ρ(M) ≥ 0 the spectral radius of the matrix M := I −R −1 D, i.e., the largest<br />

eigenvalue modulus among the, possibly complex, eigenvalues of M. Then we have a<br />

classical convergence result (see for instance isaacson and keller(1966), p.63).<br />

Theorem 7.6.1 - If M adm<strong>it</strong>s a diagonal form and ρ(M) < 1, then we have<br />

limk→+∞ ¯p (k) = ¯p, where ¯p is the solution to (7.6.1).<br />

The closer ρ(M) is to zero, the faster the convergence will be, since <strong>it</strong> depends on the<br />

geometric progression [ρ(M)] k , k ∈ N. Actually, when R = D, we obtain the exact<br />

solution in one step.<br />

Many famous schemes have their root in formula (7.6.2). The simplest is the Richardson<br />

method (see richardson(1910)) obtained by setting R −1 = θI, where θ ∈ R . Let us<br />

indicate by λm ∈ C, 1 ≤ m ≤ n, the eigenvalues of D. If D adm<strong>it</strong>s a diagonal form<br />

and the following hypotheses are satisfied:<br />

(7.6.3) Reλm > 0 and 0 < θ <<br />

2 Reλm<br />

, 1 ≤ m ≤ n,<br />

|λm| 2<br />

then ρ(M) < 1, and the Richardson scheme is convergent.<br />

To proceed w<strong>it</strong>h our investigation, we need to know more about the eigenvalues of<br />

the matrices introduced in this chapter. This analysis is given in chapter eight.<br />

Desp<strong>it</strong>e <strong>it</strong>s simplic<strong>it</strong>y, the Richardson scheme is useful to help understand the basic<br />

principles of <strong>it</strong>erative methods. Consequently, more elaborate <strong>it</strong>erative algor<strong>it</strong>hms will<br />

not be discussed.


8<br />

EIGENVALUE ANALYSIS<br />

The behavior of the eigenvalues of the matrices introduced in chapter seven is interesting<br />

for two reasons. The first is related to the application of <strong>it</strong>erative methods for the<br />

solution of the corresponding systems of equations. The second is to examine the<br />

approximation of eigenvalue problems for differential operators, which arise in various<br />

applications in mathematical physics.<br />

8.1 Eigenvalues of first derivative operators<br />

All the eigenvalues of the derivative matrices presented in sections 7.1 and 7.2, are equal<br />

to zero. Actually, if there existed one eigenvalue different from zero, a corresponding<br />

polynomial eigenfunction would not vanish after repeated applications of the derivative<br />

operator. This is in contradiction to one of the fundamental theorems of calculus.<br />

The s<strong>it</strong>uation changes when we introduce auxiliary constraints, such as boundary con-<br />

d<strong>it</strong>ions. Now, since the matrix is required to be non singular, all the corresponding<br />

eigenvalues must be different from zero.<br />

Let us begin by analyzing the eigenvalues of the matrix associated w<strong>it</strong>h problem<br />

(7.4.1), relative to the Jacobi case w<strong>it</strong>h n ≥ 1, α > −1 and β > −1. More precisely,


152 Polynomial Approximation of Differential Equations<br />

we are concerned w<strong>it</strong>h finding n + 1 complex polynomials pn,m, 0 ≤ m ≤ n, of degree<br />

at most n, and n + 1 complex numbers λn,m, 0 ≤ m ≤ n, such that<br />

(8.1.1)<br />

⎧<br />

⎨<br />

⎩<br />

p ′ n,m(η (n)<br />

i ) = λn,m pn,m(η (n)<br />

i ), 1 ≤ i ≤ n,<br />

pn,m(η (n)<br />

0 ) = λn,m pn,m(η (n)<br />

0 ),<br />

0 ≤ m ≤ n.<br />

Each pn,m, 0 ≤ m ≤ n (different from the zero function) is determined up to a constant<br />

factor. Let us suppose that pn,m(η (n)<br />

0 ) = 0, for m = 0. Then, <strong>it</strong> is easy to realize that<br />

λn,0 = 1. On the contrary, the remaining λn,m, 1 ≤ m ≤ n, are eigenvalues of the<br />

n × n matrix, obtained by assuming σ = 0 in (7.4.1) (in (7.4.3) we have an example<br />

for n = 3). Therefore, except for m = 0, the eigenfunctions satisfy pn,m(η (n)<br />

0 ) = 0.<br />

In figures 8.1.1 to 8.1.4, we give the plot of the eigenvalues λn,m, 1 ≤ m ≤ n, in the<br />

window [−4,12] × [−12,12] of the complex plane. We considered respectively the four<br />

cases: α = β = −.7, −.5, 0, .5 . The integer n varies from 4 to 10. The eigenvalues,<br />

corresponding to the same value of n, have been joined together by segments.<br />

In the first three cases, all the eigenvalues lie in the half plane of the complex numbers<br />

w<strong>it</strong>h pos<strong>it</strong>ive real part. In general, we are induced to conjecture that Reλn,m > 0,<br />

1 ≤ m ≤ n, when α > −1, −1 < β ≤ 0. Unfortunately, we do not have the proof of<br />

this fact. The statement has been proven in solomonoff and turkel(1989), for the<br />

Chebyshev case (α = β = −.5), and could be adapted to other s<strong>it</strong>uations. Anyway, the<br />

proof is qu<strong>it</strong>e technical and we om<strong>it</strong> <strong>it</strong>.<br />

We further characterize the eigenvalues w<strong>it</strong>h the following propos<strong>it</strong>ion.<br />

Theorem 8.1.1 - For 1 ≤ m ≤ n, λn,m satisfies the relation<br />

(8.1.2) λ n n,m +<br />

n<br />

j d<br />

j=1<br />

dx j ˜ l (n)<br />

0 (−1)<br />

where the Lagrange polynomial ˜ l (n)<br />

0 is defined in (3.2.8).<br />

<br />

λ n−j<br />

n,m = 0,


Eigenvalue Analysis 153<br />

Figure 8.1.1 - Eigenvalues {λn,m}1≤m≤n Figure 8.1.2 - Eigenvalues {λn,m}1≤m≤n<br />

for 4 ≤ n ≤ 10 and α = β = −.7 . for 4 ≤ n ≤ 10 and α = β = −.5 .<br />

Figure 8.1.3 - Eigenvalues {λn,m}1≤m≤n Figure 8.1.4 - Eigenvalues {λn,m}1≤m≤n<br />

for 4 ≤ n ≤ 10 and α = β = 0 . for 4 ≤ n ≤ 10 and α = β = .5 .


154 Polynomial Approximation of Differential Equations<br />

Proof - The eigenfunctions are solutions of the equation<br />

(8.1.3) p ′ n,m = λn,mpn,m + p ′ n,m(−1) ˜l (n)<br />

0 , in Ī = [−1,1], 1 ≤ m ≤ n.<br />

By differentiating the above expression k times, we get by recursion<br />

(8.1.4)<br />

n,mpn,m + p ′ ⎡<br />

n,m(−1) ⎣λ k n,m ˜l (n)<br />

0 +<br />

k<br />

j d<br />

d k+1<br />

dx k+1 pn,m = λ k+1<br />

j=1<br />

dxj ˜l (n)<br />

0<br />

<br />

λ k−j⎦,<br />

n,m<br />

1 ≤ m ≤ n, k ≥ 1.<br />

Recalling that the derivative of order n + 1 of a polynomial of degree n is zero, and<br />

pn,m(−1) = 0, 1 ≤ m ≤ n, we take k = n in (8.1.4) and evaluate at the point x = −1.<br />

This yields (8.1.2) (<strong>it</strong> is evident from (8.1.3) that p ′ n,m(−1) = 0, 1 ≤ m ≤ n).<br />

In other words, the λn,m’s are the roots of the characteristic polynomial of degree n<br />

associated w<strong>it</strong>h our eigenvalue problem. By directly solving the differential equation in<br />

(8.1.3), we obtain the following explic<strong>it</strong> expression for the eigenfunctions:<br />

(8.1.5) pn,m(x) = p ′ n,m(−1) e λn,mx<br />

x<br />

−1<br />

e −λn,mt ˜(n) l 0 (t) dt, x ∈ Ī, 1 ≤ m ≤ n.<br />

Some properties can be deduced from formulas (8.1.2) and (8.1.5), but the most inter-<br />

esting conjectures follow from numerical experiments. We will give other specifications<br />

in section 8.3. We must pay attention when we carry out tests on problem (8.1.1). In-<br />

deed, in trefethen and trummer(1987), the authors note that disagreeable rounding<br />

errors occur when computing the eigenvalues w<strong>it</strong>hout appropriate machine accuracy.<br />

Conclusions from careless computations could sometimes be misleading.<br />

(7.4.4) is<br />

(8.1.6)<br />

Let us proceed w<strong>it</strong>h our investigation. The eigenvalue problem associated w<strong>it</strong>h<br />

⎧<br />

⎨<br />

⎩<br />

p ′ n,m(η (n)<br />

i ) = λn,m pn,m(η (n)<br />

i ), 1 ≤ i ≤ n,<br />

p ′ n,m(η (n)<br />

0 ) + γpn,m(η (n)<br />

0 ) = λn,m pn,m(η (n)<br />

0 ),<br />

⎤<br />

0 ≤ m ≤ n.


Eigenvalue Analysis 155<br />

We find from numerical tests that the λn,m’s have pos<strong>it</strong>ive real part when −1 < β ≤ 0<br />

and γ > γn, where γn, n ≥ 1, is a pos<strong>it</strong>ive real constant. This result is found in<br />

funaro and gottlieb(1988) for α = β = −1/2. Here, we give a simplified version in<br />

the Legendre case.<br />

Theorem 8.1.2 - Let α = β = 0. Then, for any n ≥ 1, there exists a constant<br />

γn > 0, such that, for γ > γn, all the eigenvalues of problem (8.1.6) satisfy Reλn,m > 0,<br />

0 ≤ m ≤ n.<br />

Proof - We first note that pn,m(η (n)<br />

0 ) = 0, 0 ≤ m ≤ n. Otherwise, we would have the<br />

differential equation p ′ n,m = λn,mpn,m which is satisfied only for pn,m ≡ 0. Then, for<br />

0 ≤ m ≤ n, we get the relations (the upper bar denotes the complex conjugate)<br />

(8.1.7)<br />

d<br />

dx (|pn,m| 2 )(η (n)<br />

i ) = (pn,m¯p ′ n,m + p ′ n,m¯pn,m)(η (n)<br />

i )<br />

= ( ¯ λn,mpn,m¯pn,m + λn,mpn,m¯pn,m)(η (n)<br />

i ) = 2[|pn,m|(η (n)<br />

i )] 2 Reλn,m, 1 ≤ i ≤ n,<br />

(8.1.8) 2γ[|pn,m|(η (n)<br />

0 )] 2 + d<br />

dx (|pn,m| 2 )(η (n)<br />

0 ) = 2[|pn,m|(η (n)<br />

0 )] 2 Reλn,m.<br />

We multiply the expressions in (8.1.7), (8.1.8) by the weights ˜w (n)<br />

i , 0 ≤ i ≤ n, given in<br />

(3.5.6). By summing over i, one obtains<br />

(8.1.9) 2γ[|pn,m|(η (n)<br />

0 )] 2 ˜w (n)<br />

0 +<br />

= 2Reλn,m<br />

n<br />

i=0<br />

n<br />

i=0<br />

d<br />

dx (|pn,m| 2 )(η (n)<br />

i ) ˜w (n)<br />

i<br />

[|pn,m|(η (n)<br />

i )] 2 ˜w (n)<br />

i , 0 ≤ m ≤ n.<br />

Since the weights are pos<strong>it</strong>ive, <strong>it</strong> is sufficient to show that the term on the left-hand<br />

side of (8.1.9) is also pos<strong>it</strong>ive. By virtue of formula (3.5.1) and theorem 3.5.1 (w ≡ 1),<br />

considering that d<br />

dx (|pn,m| 2 ) ∈ P2n−1, we get


156 Polynomial Approximation of Differential Equations<br />

(8.1.10) 2γ[pn,m|(η (n)<br />

0 )] 2 ˜w (n)<br />

0 +<br />

= (2γ ˜w (n)<br />

0<br />

By taking γn := (2 ˜w (n)<br />

0 ) −1 = 1<br />

4<br />

n<br />

i=0<br />

= 2γ[|pn,m|(η (n)<br />

0 )] 2 ˜w (n)<br />

0 +<br />

d<br />

dx (|pn,m| 2 )(η (n)<br />

i ) ˜w (n)<br />

i<br />

1<br />

−1<br />

d<br />

dx |pn,m| 2 dx<br />

− 1)[|pn,m|(η (n)<br />

0 )] 2 + [|pn,m|(η (n)<br />

n )] 2 , 0 ≤ m ≤ n.<br />

n(n + 1), we conclude the proof.<br />

We note that γn must be proportional to n 2 , to ensure that the eigenvalues have<br />

pos<strong>it</strong>ive real part. This is also true for the Chebyshev case. In general, there is a real<br />

eigenvalue growing linearly w<strong>it</strong>h γ. When γ tends to infin<strong>it</strong>y, this eigenvalue diverges<br />

and the remaining n eigenvalues approach those of problem (8.1.1) for 1 ≤ m ≤ n.<br />

Heuristically, for γ = +∞, we are forcing the eigenfunctions to satisfy a vanishing<br />

boundary cond<strong>it</strong>ion at the point x = −1.<br />

Setting q = λp in (7.4.5) and arguing as in theorem 8.1.1, we obtain, for any γ ∈ R<br />

n<br />

<br />

j d<br />

(8.1.11) λ n+1<br />

n,m + γλ n n,m + γ<br />

j=1<br />

dx j ˜ l (n)<br />

0 (−1)<br />

λ n−j<br />

n,m = 0, 0 ≤ m ≤ n.<br />

For boundary cond<strong>it</strong>ions at x = 1, similar statements hold. Now, the counterparts<br />

of (8.1.1) and (8.1.6) are respectively the following ones:<br />

⎧<br />

⎨<br />

(8.1.12)<br />

⎩<br />

(8.1.13)<br />

⎧<br />

⎨<br />

⎩<br />

−p ′ n,m(η (n)<br />

i ) = λn,m pn,m(η (n)<br />

i ), 1 ≤ i ≤ n,<br />

pn,m(η (n)<br />

n ) = λn,m pn,m(η (n)<br />

n ),<br />

−p ′ n,m(η (n)<br />

i ) = λn,m pn,m(η (n)<br />

i ), 1 ≤ i ≤ n,<br />

−p ′ n,m(η (n)<br />

n ) + γpn,m(η (n)<br />

n ) = λn,m pn,m(η (n)<br />

n ),<br />

0 ≤ m ≤ n.<br />

The change in the sign ensures that the eigenvalues verify Reλn,m > 0, 0 ≤ m ≤ n,<br />

provided that −1 < α ≤ 0, β > −1, and γ > γn > 0.


Eigenvalue Analysis 157<br />

Finally, the analysis of the eigenvalue problems associated to (7.4.7) and (7.4.8)<br />

leads to pos<strong>it</strong>ive real parts when α > −1, −1 < β ≤ 0, γ > γn > 0, and when A is a<br />

non negative function.<br />

8.2 Eigenvalues of higher-order operators<br />

Let us begin w<strong>it</strong>h the discussion of second-order problems in the Jacobi case. The<br />

eigenvalue problem corresponding to (7.4.9) consists in finding n+1 complex polynomials<br />

pn,m, 0 ≤ m ≤ n, of degree at most n (n ≥ 2), and n + 1 complex numbers λn,m,<br />

0 ≤ m ≤ n, such that<br />

⎧<br />

(8.2.1)<br />

⎪⎨<br />

⎪⎩<br />

−p ′′ n,m(η (n)<br />

i ) = λn,mpn,m(η (n)<br />

i ), 1 ≤ i ≤ n − 1,<br />

pn,m(η (n)<br />

0 ) = λn,mpn,m(η (n)<br />

0 ),<br />

pn,m(η (n)<br />

n ) = λn,mpn,m(η (n)<br />

n ),<br />

0 ≤ m ≤ n.<br />

We assume that pn,0(η (n)<br />

0 ) = 0 and pn,n(η (n)<br />

n ) = 0. This implies that λn,0 = λn,n = 1.<br />

The other n − 1 eigenvalues are obtained from the reduced system (7.4.9) after setting<br />

σ1 = σ2 = 0 (see for instance (7.4.11), corresponding to the case n = 3). The relative<br />

eigenfunctions satisfy pn,m(η (n)<br />

0 ) = pn,m(η (n)<br />

n ) = 0, 1 ≤ m ≤ n − 1.<br />

To construct the characteristic polynomial, we argue as in theorem 8.1.1. We briefly<br />

outline the proof. This time, we start from the relation<br />

(8.2.2) −p ′′ n,m = λn,mpn,m + p ′′ n,m(−1) ˜ l (n)<br />

0 + p′′ n,m(1) ˜ l (n)<br />

n , 1 ≤ m ≤ n − 1.<br />

We differentiate both the terms of equal<strong>it</strong>y (8.2.2), until the left-hand side vanishes.<br />

Evaluating the resulting expression at the points x = ±1, leads to a 2 × 2 linear<br />

system in the unknowns p ′′ n,m(−1) and p ′′ n,m(1). The eigenfunction pn,m in (8.2.2)<br />

is not uniquely determined, since <strong>it</strong> can differ by a multiplicative constant. Hence, the<br />

determinant of the 2 × 2 system must vanish. Imposing this final cond<strong>it</strong>ion yields the<br />

characteristic polynomial.


158 Polynomial Approximation of Differential Equations<br />

Also for problem (8.2.1), we have eigenvalues w<strong>it</strong>h pos<strong>it</strong>ive real part. We first prove<br />

a preliminary result in the ultraspherical case.<br />

Lemma 8.2.1 - Let ν := α = β w<strong>it</strong>h −1 < ν ≤ 1. Then we can find a constant<br />

C > 0 such that, for any n ≥ 2 and p ∈ Pn satisfying p(±1) = 0, one has<br />

(8.2.3)<br />

1<br />

p<br />

−1<br />

′ (pw) ′ 1<br />

dx ≥ C [p<br />

−1<br />

′ ] 2 w dx.<br />

Proof - We follow the guideline of the proof given in canuto and quarteroni(1981),<br />

in the case ν = −1/2. First we have<br />

(8.2.4)<br />

1<br />

p<br />

−1<br />

′ (pw) ′ dx =<br />

1<br />

−1<br />

[p ′ ] 2 w dx − 1<br />

1<br />

p<br />

2 −1<br />

2 w ′′ dx,<br />

where we integrated by parts, using the cond<strong>it</strong>ion p(±1) = 0. If 0 ≤ ν ≤ 1, the proof<br />

is complete w<strong>it</strong>h C = 1, since w ′′ ≤ 0. On the other hand, let −1 < ν < 0 and define<br />

τ := ν−1<br />

2ν<br />

(8.2.5)<br />

+ τ 2<br />

=<br />

> 0. Then, we have<br />

1<br />

Considering that<br />

1<br />

p<br />

−1<br />

′ (pw) ′ dx =<br />

−1<br />

1<br />

(p<br />

−1<br />

′ w + τpw ′ ) 2 w −1 dx +<br />

1<br />

[p<br />

−1<br />

′ ] 2 1<br />

w dx + 2τ pp<br />

−1<br />

′ w ′ dx<br />

p 2 (w ′ ) 2 w −1 dx + ( 1<br />

1<br />

2 − τ) [p<br />

−1<br />

2 ] ′ w ′ dx − τ 2<br />

1<br />

p<br />

−1<br />

2 (w ′ ) 2 w −1 dx<br />

1<br />

(8.2.6) (τ − 1<br />

2 )w′′ − τ 2 (w ′ ) 2 w −1 ≥ −<br />

formula (8.2.5) yields<br />

(8.2.7)<br />

−1<br />

1<br />

p<br />

−1<br />

′ (pw) ′ dx ≥ −<br />

p 2 (τ − 1<br />

2 )w′′ − τ 2 (w ′ ) 2 w −1 dx.<br />

1 + ν<br />

4ν<br />

1 + ν<br />

4ν w′′ ,<br />

1<br />

p<br />

−1<br />

2 w ′′ dx.


Eigenvalue Analysis 159<br />

Combining w<strong>it</strong>h (8.2.4), we finally get<br />

(8.2.8)<br />

1<br />

p<br />

−1<br />

′ (pw) ′ dx ≥<br />

1<br />

−1<br />

Inequal<strong>it</strong>y (8.2.3) follows w<strong>it</strong>h C = 1+ν<br />

1−ν .<br />

[p ′ ] 2 w dx + 2ν<br />

1<br />

p<br />

1 + ν −1<br />

′ (pw) ′ dx.<br />

Another proof of lemma 8.2.1 is given in bernardi and maday(1989).<br />

When the polynomials are complex, the proof of lemma 8.2.1 is identical to that given<br />

above. In this way, we can find a constant C > 0, such that<br />

1<br />

(8.2.9) Re p ′ (¯pw) ′ 1<br />

dx ≥ C<br />

−1<br />

We can now state the next propos<strong>it</strong>ion.<br />

|p<br />

−1<br />

′ | 2 w dx.<br />

Theorem 8.2.2 - Let ν := α = β w<strong>it</strong>h −1 < ν ≤ 1. Then, for any n ≥ 2, the<br />

eigenvalues of problem (8.2.1) satisfy Reλn,m > 0, 1 ≤ m ≤ n − 1.<br />

Proof - We remark that formula (3.5.1) and theorem 3.5.1 also apply to complex<br />

polynomials. It suffices to argue w<strong>it</strong>h the real and imaginary parts separately.<br />

Recalling that pn,m(±1) = 0, we use integration by parts to obtain<br />

(8.2.10)<br />

= −<br />

n<br />

i=0<br />

1<br />

−1<br />

p ′′ n,m(η (n)<br />

i )¯pn,m(η (n)<br />

i ) ˜w (n)<br />

i<br />

p ′ n,m(¯pn,mw) ′ 1<br />

dx = − p<br />

−1<br />

′′ n,m¯pn,mw dx<br />

= λn,m<br />

n<br />

i=0<br />

[|pn,m|(η (n)<br />

i )] 2 ˜w (n)<br />

i , 1 ≤ m ≤ n − 1.<br />

We first note that p ′ n,m ≡ 0 for all m. Otherwise, the vanishing cond<strong>it</strong>ions at the<br />

boundaries would imply pn,m ≡ 0 for some m. Recalling (8.2.9), we easily conclude<br />

the proof, since the real part of the left-hand side in (8.2.10) is strictly pos<strong>it</strong>ive.<br />

A straightforward and meaningful corollary is obtained by examining relation (8.2.10).


160 Polynomial Approximation of Differential Equations<br />

Theorem 8.2.3 - Let α = β = 0. Then, for any n ≥ 2, the eigenvalues of problem<br />

(8.2.1) are real and strictly pos<strong>it</strong>ive.<br />

Proof - Given that w ≡ 1, <strong>it</strong> is sufficient to recall the equal<strong>it</strong>y<br />

(8.2.11)<br />

1<br />

p<br />

−1<br />

′ n,m(¯pn,mw) ′ dx =<br />

1<br />

|p<br />

−1<br />

′ n,m| 2 dx > 0, 1 ≤ m ≤ n − 1.<br />

For different values of the parameters α and β, problem (8.2.1) still furnishes real<br />

eigenvalues, though the proof of this statement is more involved than for the Legendre<br />

case. By studying the roots of the characteristic polynomial, gottlieb and lust-<br />

man(1983) give the proof in the Chebyshev case. To this end, they analyze the eigen-<br />

values of two auxiliary first-order problems. Other cases are still an open question. For<br />

−1 < α < 1 and −1 < β < 1 , the numerical experiments generate real eigenvalues,<br />

and some complex eigenvalues w<strong>it</strong>h relatively small imaginary part. In general, the<br />

analysis of the eigenvalue problem related to (7.4.12) does not lead to real eigenvalues,<br />

due to the presence of the first-order derivative (see previous section). However, when<br />

A and B are sufficiently close to zero, the second order operator dominates, and the<br />

eigenvalues are a small perturbation of those in (8.2.1). Thus, their real part is pos<strong>it</strong>ive.<br />

We examine the eigenvalue problems associated to (7.4.13) and (7.4.16). Namely,<br />

for 0 ≤ m ≤ n, we have<br />

⎧<br />

(8.2.12)<br />

(8.2.13)<br />

⎪⎨<br />

⎪⎩<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

−p ′′ n,m(η (n)<br />

i ) + µpn,m(η (n)<br />

i ) = λn,mpn,m(η (n)<br />

i ), 1 ≤ i ≤ n − 1,<br />

−p ′ n,m(η (n)<br />

0 ) = λn,mpn,m(η (n)<br />

0 ),<br />

p ′ n,m(η (n)<br />

n ) = λn,mpn,m(η (n)<br />

n ),<br />

−p ′′ n,m(η (n)<br />

i ) + µpn,m(η (n)<br />

i ) = λn,mpn,m(η (n)<br />

i ), 1 ≤ i ≤ n − 1,<br />

−p ′′ n,m(η (n)<br />

0 ) + µpn,m(η (n)<br />

0 ) − γp ′ n,m(η (n)<br />

0 ) = λn,mpn,m(η (n)<br />

0 ),<br />

−p ′′ n,m(η (n)<br />

n ) + µpn,m(η (n)<br />

n ) + γp ′ n,m(η (n)<br />

n ) = λn,mpn,m(η (n)<br />

n ),<br />

where µ > 0 and γ ∈ R w<strong>it</strong>h γ = 0.


Eigenvalue Analysis 161<br />

We note that constant polynomials are eigenfunctions of problem (8.2.13), corresponding<br />

to the eigenvalue µ. Very l<strong>it</strong>tle theory has been developed. We only present the following<br />

result.<br />

Theorem 8.2.4 - Let α = β = 0. Then, for any n ≥ 2, if γ = 1<br />

2n(n + 1), the<br />

eigenvalues of problem (8.2.13) are real and strictly pos<strong>it</strong>ive.<br />

Proof - Noting that γ ˜w (n)<br />

0<br />

(8.2.14)<br />

= γ ˜w(n)<br />

n = 1 (see (3.5.6)), we get<br />

1<br />

|p<br />

−1<br />

′ n,m| 2 n<br />

dx + µ<br />

i=0<br />

= [p ′ n,m¯pn,m](1) − [p ′ n,m¯pn,m](−1) −<br />

= γ ˜w (n)<br />

n [p ′ n,m¯pn,m](1) − γ ˜w (n)<br />

0 [p ′ n,m¯pn,m](−1) +<br />

The proof follows easily.<br />

= λn,m<br />

n<br />

i=0<br />

[|pn,m|(η (n)<br />

i )] 2 ˜w (n)<br />

i<br />

1<br />

p<br />

−1<br />

′′ n<br />

n,m¯pn,m dx + µ [|pn,m|(η<br />

i=0<br />

(n)<br />

i )] 2 ˜w (n)<br />

i<br />

n<br />

i=0<br />

[(−p ′′ n,m + µpn,m)¯pn,m](η (n)<br />

i ) ˜w (n)<br />

i<br />

[|pn,m|(η (n)<br />

i )] 2 ˜w (n)<br />

i , 0 ≤ m ≤ n.<br />

Real or complex eigenvalues, w<strong>it</strong>h a relatively small imaginary part, are obtained for<br />

many other values of the parameters α and β, both for problems (8.2.12) and (8.2.13).<br />

In the latter case, γ is required to be pos<strong>it</strong>ive and larger than a constant γn, which<br />

is in general proportional to n 2 . To our knowledge, currently, there are no theoretical<br />

results to support these conjectures.<br />

It is worthwhile to mention that an incorrect specification of the sign of the equa-<br />

tion, when imposing boundary cond<strong>it</strong>ions (see for instance (7.4.13)), can result in a<br />

matrix w<strong>it</strong>h at least one negative eigenvalue. For example, replacing the second equa-<br />

tion in (8.2.12) by p ′ n,m(η (n)<br />

0 ) = λn,mpn,m(η (n)<br />

0 ), we still obtain real eigenvalues, but<br />

one of them is negative.


162 Polynomial Approximation of Differential Equations<br />

Let us investigate the case of fourth-order operators. Here, λn,m, 1 ≤ m ≤ n − 1,<br />

and rn,m, 1 ≤ m ≤ n − 1, are respectively the eigenvalues and the eigenfunctions<br />

associated to problem (7.4.22). Therefore, we can wr<strong>it</strong>e r IV<br />

n,m(η (n)<br />

i ) = λn,mrn,m(η (n)<br />

i ),<br />

1 ≤ i ≤ n − 1, and rn,m(±1) = r ′ n,m(±1) = 0.<br />

Before proceeding w<strong>it</strong>h our analysis, we state the following propos<strong>it</strong>ion.<br />

Lemma 8.2.5 - Let ν := α = β w<strong>it</strong>h −1 < ν ≤ 1. Then, we can find a constant<br />

C > 0, such that, for any n ≥ 2 and p ∈ Pn+2 satisfying p(±1) = p ′ (±1) = 0, one<br />

has<br />

(8.2.15)<br />

1<br />

p<br />

−1<br />

′′ (pw) ′′ 1<br />

dx ≥ C [p<br />

−1<br />

′′ ] 2 w dx.<br />

The proof of the above result can be found in bernardi and maday(1990). It is qu<strong>it</strong>e<br />

technical, but very similar to that of theorem 8.2.1. The same proof, adapted to the<br />

case of complex polynomials, yields the inequal<strong>it</strong>y<br />

1<br />

(8.2.16) Re p ′′ (¯pw) ′′ <br />

dx ≥ C<br />

We are now ready to prove the following propos<strong>it</strong>ion.<br />

−1<br />

1<br />

|p<br />

−1<br />

′′ | 2 w dx.<br />

Theorem 8.2.6 - Let ν := α = β w<strong>it</strong>h −1 < ν ≤ 1. Then, for any n ≥ 2,<br />

the eigenvalues relative to the system (7.4.22) satisfy Reλn,m > 0, 1 ≤ m ≤ n − 1.<br />

Moreover, when ν = 0, we have λn,m ∈ R, 1 ≤ m ≤ n − 1.<br />

Proof - One has<br />

n<br />

(8.2.17)<br />

i=0<br />

r IV<br />

n,m(η (n)<br />

i )¯rn,m(η (n)<br />

i ) ˜w (n)<br />

i<br />

= λn,m<br />

n<br />

i=0<br />

[|rn,m|(η (n)<br />

i )] 2 ˜w (n)<br />

i .<br />

We must show that the left-hand side of (8.2.17) has pos<strong>it</strong>ive real part. Now, for<br />

1 ≤ m ≤ n − 1, we can find a complex number ζn,m ∈ C such that<br />

(8.2.18)<br />

1<br />

1 − x 2 rn,m(x) = ζn,mun(x) + {lower degree terms},


Eigenvalue Analysis 163<br />

(8.2.19) (1 − x 2 ) r IV<br />

n,m(x) = n(n 2 − 1)(n + 2)ζn,mun(x) + {lower degree terms},<br />

where un = P (ν,ν)<br />

n . These relations can be checked by equating the coefficients of the<br />

monomials x n . W<strong>it</strong>h the same arguments used in section 3.8 (we recall for instance<br />

(3.8.12) and (3.8.15)), <strong>it</strong> is easy to show that<br />

(8.2.20)<br />

=<br />

1<br />

−1<br />

n<br />

i=0<br />

r IV<br />

n,m(η (n)<br />

i )¯rn,m(η (n)<br />

i ) ˜w (n)<br />

i<br />

r IV<br />

n,m¯rn,m w dx + (un 2 w,n − un 2 w) n(n 2 − 1)(n + 2)|ζn,m| 2 .<br />

Having unw,n ≥ unw, ∀n ≥ 1 (see (2.2.10) and (3.8.3)), integration by parts and<br />

formula (8.2.16) yield<br />

(8.2.21) Re<br />

n<br />

i=0<br />

1<br />

= Re r<br />

−1<br />

′′<br />

n,m(¯rn,mw) ′′ <br />

dx<br />

r IV<br />

n,m(η (n)<br />

i )¯rn,m(η (n)<br />

i ) ˜w (n)<br />

i<br />

≥ C<br />

1<br />

−1<br />

<br />

1<br />

≥ Re r<br />

−1<br />

IV<br />

<br />

n,m¯rn,mw dx<br />

|r ′′<br />

n,m| 2 w dx > 0, 1 ≤ m ≤ n − 1.<br />

We note that, since rn,m(±1) = r ′ n,m(±1) = 0, the last integral in (8.2.21) does not<br />

vanish (otherwise we would have rn,m ≡ 0). Hence, by (8.2.17) we deduce Reλn,m > 0,<br />

1 ≤ m ≤ n − 1.<br />

When ν = 0 (thus w ≡ 1), the right-hand side in (8.2.20) turns out to be real, after<br />

integrating by parts twice.<br />

Real and pos<strong>it</strong>ive eigenvalues are also observed for other values of the parameters<br />

α and β, and also for different type of boundary cond<strong>it</strong>ions such as those in (7.4.23).<br />

Hints for the determination of the characteristic polynomial are given in funaro and<br />

heinrichs(1990).


164 Polynomial Approximation of Differential Equations<br />

Similar arguments hold for the case of unbounded domains. Let us see some exam-<br />

ple. Here, we denote by λn,m and pn,m, 0 ≤ m ≤ n − 1, respectively the eigenvalues<br />

and the eigenfunctions related to problem (7.4.26). In particular, the pn,m’s are com-<br />

plex polynomials of degree at most n − 1. Again, we have λn,0 = 1 and pn,m(0) = 0,<br />

for 1 ≤ m ≤ n − 1.<br />

We recall that the space of Laguerre functions Sn has been introduced in section 6.7.<br />

As in the previous cases, we need a preparatory result.<br />

Lemma 8.2.7 - Let −1 < α < 1 and v(x) := x α e x , x ∈]0,+∞[. Then, we can find<br />

a constant C > 0 such that, for any n ≥ 1 and P ∈ Sn−1, satisfying P(0) = 0, one<br />

has<br />

(8.2.22)<br />

+∞<br />

where µ > 1<br />

2 max{1, 1<br />

1−α }.<br />

0<br />

P ′ (Pv) ′ +∞<br />

dx + µ P<br />

0<br />

2 v dx ≥ C<br />

+∞<br />

[(P<br />

0<br />

′ ) 2 + P 2 ] v dx,<br />

For the proof we refer to kavian(1990). Combining this lemma w<strong>it</strong>h (7.2.25) and the<br />

experience we have gained in the previous examples, we obtain the following result.<br />

Theorem 8.2.8 - Let −1 < α < 1 and µ > 1<br />

2 max{1, 1<br />

1−α }. Then, for any n ≥ 1,<br />

the eigenvalues relative to the system (7.4.26) satisfy Reλn,m > 0, 0 ≤ m ≤ n − 1.<br />

Actually, in the Laguerre case, the eigenvalues of the operator considered are real.<br />

Nevertheless, there is no proof of this.<br />

Concerning problem (7.5.4), where the unknowns are the point values of a scaled<br />

function, the eigenvalues are identical to those of problem (7.4.26). Actually, defin<strong>it</strong>ion<br />

(7.5.1) corresponds to a linear change of variables which does not alter the spectrum of<br />

the derivative matrix.<br />

For the matrix in (7.2.15), w<strong>it</strong>h the vanishing boundary cond<strong>it</strong>ion at the point<br />

x = 0, the eigenvalues are complex, but their real part is pos<strong>it</strong>ive provided that α ≤ 0.


Eigenvalue Analysis 165<br />

In particular, when α = 0, the real part of these eigenvalues is constantly equal to 1<br />

2 .<br />

The proof of this fact is qu<strong>it</strong>e simple and is left as exercise.<br />

Finally, we consider problem (7.4.28) in which the eigenvalues and eigenfunctions<br />

are respectively denoted by λn,m and pn,m, 1 ≤ m ≤ n. This time, the theory is simple.<br />

Comparing (7.4.28) w<strong>it</strong>h the differential equation (1.7.1), we get λn,m = 2(m − 1) + µ<br />

and pn,m = Hm−1, 1 ≤ m ≤ n. Therefore, the eigenvalues are real and pos<strong>it</strong>ive when<br />

µ > 0.<br />

8.3 Cond<strong>it</strong>ion number<br />

The cond<strong>it</strong>ion number is a useful quant<strong>it</strong>y to estimate the speed of convergence of an<br />

<strong>it</strong>erative method for the solution of linear systems.<br />

It is well-known that a real n ×n matrix D corresponds to a linear operator from R n<br />

to R n . Denoting by L(R n ,R n ) the space of linear mappings from R n in R n , we can<br />

define the norm of D (see for instance isaacson and keller(1966), p.37) as follows:<br />

(8.3.1) D L(R n ,R n ) := sup<br />

¯x∈R n<br />

¯x=0<br />

D¯x<br />

¯x<br />

= sup<br />

¯x∈R n<br />

¯x=1<br />

D¯x,<br />

where by · we indicate the canonical norm in R n . The expression in (8.3.1) actually<br />

satisfies all the properties required in section 2.1. In practice, the norm of D measures<br />

the maximal deformation, when applying the matrix to the vectors of the un<strong>it</strong>ary sphere<br />

of R n .<br />

We remark that, for any n × n matrix D and any ¯x ∈ R n , one has the inequal<strong>it</strong>y<br />

(8.3.2) D¯x ≤ D L(R n ,R n )¯x.<br />

At this point, we can define the cond<strong>it</strong>ion number of a nonsingular matrix D. This is<br />

the real pos<strong>it</strong>ive number given by<br />

(8.3.3) κ(D) := D L(R n ,R n )D −1 L(R n ,R n ),


166 Polynomial Approximation of Differential Equations<br />

from which one can easily deduce κ(D) ≥ 1.<br />

This quant<strong>it</strong>y gives an idea of the distribution of the eigenvalues of D in the complex<br />

plane. When κ(D) is large, we generally expect scattered eigenvalues w<strong>it</strong>h considerable<br />

variations in their magn<strong>it</strong>ude. On the other hand, when κ(D) is close to 1, the moduli<br />

of the eigenvalues are gathered together in a small interval. We can be more precise for<br />

symmetric matrices. In this s<strong>it</strong>uation, we have the relation<br />

(8.3.4) κ(D) = π(D) := max1≤m≤n |λm|<br />

min1≤m≤n |λm| ,<br />

where the λm’s are the eigenvalues of D.<br />

When D is not symmetric we have the inequal<strong>it</strong>y: 1 ≤ π(D) ≤ κ(D).<br />

Nearly all the matrices analyzed in the sequel are not symmetric. Nevertheless, for<br />

these matrices, the quant<strong>it</strong>ies π(D) and κ(D) display more or less the same behavior.<br />

So that, allowing ourselves an abuse of terminology, the two concepts will often be<br />

interchanged.<br />

To better understand the util<strong>it</strong>y of the cond<strong>it</strong>ion number, we return to the <strong>it</strong>erative<br />

method described in section 7.6. The speed of convergence of the scheme (7.6.2) depends<br />

on the spectral radius ρ(M), where M := I − R −1 D. In the Richardson method we<br />

have<br />

(8.3.5) ρ(M) = max |1 − θλm|,<br />

1≤m≤n<br />

where the λm’s are the eigenvalues of D, satisfying the cond<strong>it</strong>ions in (7.6.3).<br />

If the eigenvalues are clustered around a particular real value, then we can find θ such<br />

that ρ(M) is near zero. This will result in a fast convergence. On the contrary, if<br />

the eigenvalues are scattered, the parameter θ cannot control all of them, and ρ(M)<br />

will be dangerously close to 1. In short, we have to beware of matrices w<strong>it</strong>h a large<br />

cond<strong>it</strong>ion number. These are called ill-cond<strong>it</strong>ioned matrices.<br />

Unfortunately, all the matrices considered in section 7.4 are ill-cond<strong>it</strong>ioned. As an<br />

example, we examine the eigenvalue problem (8.1.1). Since the λn,m’s are distinct, the<br />

corresponding matrix adm<strong>it</strong>s a diagonal form. From figures 8.1.1 to 8.1.4, <strong>it</strong> is clear that


Eigenvalue Analysis 167<br />

the magn<strong>it</strong>ude of the eigenvalues grows w<strong>it</strong>h n. It is especially for the eigenvalues w<strong>it</strong>h<br />

the largest modulus that the ratio between the real part and the imaginary part is small.<br />

According to (7.6.3), this yields a very restrictive cond<strong>it</strong>ion on θ. For instance, in the<br />

Chebyshev case, the parameter θ is required to be proportional to 1/n 2 to achieve<br />

convergence when applying the Richardson method. This makes the spectral radius in<br />

(8.3.5) very close to 1. Consequently, the Richardson scheme is very ineffective. The<br />

s<strong>it</strong>uation is worse for the Legendre case.<br />

Concerning second-order or fourth-order problems, no improvements are observed.<br />

Although the eigenvalues are in general real and pos<strong>it</strong>ive, they are qu<strong>it</strong>e scattered. For<br />

instance, referring to problem (8.2.1), <strong>it</strong> is not difficult to obtain the estimate<br />

(8.3.6) c1 ≤ |λn,m| ≤ c2 n 4 , 1 ≤ m ≤ n − 1,<br />

where c1 and c2 are pos<strong>it</strong>ive constants.<br />

Moreover, assuming that the λn,m’s are real and in increasing order, we find that λn,1<br />

converges to a real number for n → +∞, while λn,n−1 grows like n 4 . Sharper<br />

estimates are given in weideman and trefethen(1988) in the Chebyshev case, and<br />

in vandeven (1990) in the Legendre case. For the proof of (8.3.6), when −1 < α < 1,<br />

−1 < β < 1, we refer the reader to section 8.6.<br />

Thus, the cond<strong>it</strong>ion number of the matrix corresponding to the system (7.4.9) is ex-<br />

pected to be very large. The minimal spectral radius obtained from (8.3.5) is ρ =<br />

(λn,n−1 − λn,1)/(λn,n−1 + λn,1), relative to the choice θ := 2/(λn,n−1 + λn,1) (see<br />

isaacson and keller(1966), p.84). Therefore, ρ is very close to 1, and the con-<br />

vergence of the scheme is unbelievably slow. The s<strong>it</strong>uation becomes extremely bad<br />

for fourth-order problems. In heinrichs(1989), the author suggests a way to replace<br />

problem (7.4.9) w<strong>it</strong>h an equivalent one, whose matrix has a reduced cond<strong>it</strong>ion number.<br />

Such an improvement is still insufficient when applied w<strong>it</strong>hin the framework of <strong>it</strong>erative<br />

techniques.<br />

It is time now to introduce the idea of precond<strong>it</strong>ioning. We note that the scheme<br />

(7.6.2) can be interpreted as a byproduct of the Richardson method w<strong>it</strong>h θ = 1, applied<br />

to the system


168 Polynomial Approximation of Differential Equations<br />

(8.3.7) (R −1 D)¯p = ¯q ∗ , where ¯q ∗ := R −1 ¯q.<br />

Of course (8.3.7) is equivalent to (7.6.1). The basic difference is that, by an appropriate<br />

choice of the matrix R, we can modify κ(R −1 D) to get a well-behaved system. Thus,<br />

we first compute the vector ¯q ∗ , then we solve (8.3.7) w<strong>it</strong>h an <strong>it</strong>erative approach. In<br />

this context, R is named precond<strong>it</strong>ioning matrix (or precond<strong>it</strong>ioner). Such a matrix<br />

must fulfill two requirements. First, R must be easily invertible, using an inexpensive<br />

algor<strong>it</strong>hm. Second, the cond<strong>it</strong>ion number of R −1 D has to be as close as possible to 1.<br />

The choice R = I (which leads to the Richardson method) satisfies the first restriction<br />

but not the latter. The choice R = D gives κ(R −1 D) = 1, but the computation of<br />

R −1 brings us back to the in<strong>it</strong>ial problem. The desired R is the right balance between<br />

these two extreme cases.<br />

The next several sections describe how to construct precond<strong>it</strong>ioners for the matrices<br />

previously analyzed.<br />

8.4 Precond<strong>it</strong>ioners for second-order operators<br />

We begin by studying problem (7.4.9). The case of first-order derivatives is more delicate<br />

and is considered later.<br />

Let D denote the (n + 1) × (n + 1) matrix corresponding to the system (7.4.9)<br />

(in (7.4.10) we have an example for n = 3). We recall that D coincides w<strong>it</strong>h the<br />

second derivative matrix − ˜ D 2 n (see section 7.2), except for two rows, which have been<br />

su<strong>it</strong>ably modified to take care of the boundary cond<strong>it</strong>ions.<br />

As remarked in section 8.3, the cond<strong>it</strong>ion number of D grows at least as n 4 . We<br />

are concerned w<strong>it</strong>h finding an appropriate precond<strong>it</strong>ioner R for D. To this end, for<br />

any n ≥ 2, we define h (n)<br />

j := η (n)<br />

j − η(n)<br />

j−1 , 1 ≤ j ≤ n, and ˆ h (n)<br />

j := 1<br />

2 (η(n)<br />

j+1 − η(n)<br />

j−1 ),<br />

1 ≤ j ≤ n − 1.


Eigenvalue Analysis 169<br />

Next, following orszag(1980), we consider the matrix R := {rij} 0≤i≤n , w<strong>it</strong>h entries<br />

0≤j≤n<br />

⎪⎨<br />

(8.4.1) rij :=<br />

⎧<br />

1 if i = j = 0 or i = j = n,<br />

−1<br />

h (n)<br />

i ˆ h (n)<br />

i<br />

2<br />

h (n)<br />

i+1h(n) i<br />

−1<br />

h (n)<br />

i+1 ˆ h (n)<br />

i<br />

⎪⎩<br />

0 elsewhere.<br />

if 1 ≤ i = j + 1 ≤ n − 1,<br />

if 1 ≤ i = j ≤ n − 1,<br />

if 1 ≤ i = j − 1 ≤ n − 1,<br />

The matrix R represents the second-order centered fin<strong>it</strong>e-differences operator defined<br />

on the grid points η (n)<br />

j , 0 ≤ j ≤ n. If f is a given regular function, the matrix R<br />

applied to the vector {f(η (n)<br />

j )}0≤j≤n furnishes an approximation of −f ′′ (η (n)<br />

i ),<br />

1 ≤ i ≤ n − 1. We remark that this kind of discretization is of local type (see section<br />

7.2), hence <strong>it</strong> does not possess the good convergence behavior of spectral methods, even<br />

if the nodes of a Gauss formula are involved. However, R is tridiagonal, thus, the<br />

computation of R −1 can be performed in a fast way. Therefore, R satisfies the first<br />

requirement to be a good precond<strong>it</strong>ioning matrix. Now, we must study the spectrum of<br />

R −1 D. We denote by Λn,m, 0 ≤ m ≤ n, the eigenvalues obtained after precond<strong>it</strong>ioning.<br />

Two of them, say Λn,0 and Λn,n, are equal to 1. Numerical experiments indicate the<br />

existence of a constant c > 1 (depending on α and β) such that, for any n ≥ 2, one<br />

has<br />

(8.4.2) 1 ≤ |Λn,m| < c, 1 ≤ m ≤ n − 1.<br />

Comparing this w<strong>it</strong>h (8.3.6), the improvement is impressive. Furthermore, the Λn,m’s<br />

are real and pos<strong>it</strong>ive and c is in general less than 2.5 . We can be more precise in the<br />

Chebyshev case. In fact, in haldenwang, labrosse, abboudi and deville(1984),


170 Polynomial Approximation of Differential Equations<br />

an exact expression is given for the precond<strong>it</strong>ioned eigenvalues. Namely, we have<br />

(8.4.3) Λn,m = m(m + 1)<br />

2 π sin 2n<br />

sin mπ<br />

2n<br />

cos π<br />

2n<br />

sin (m+1)π<br />

2n<br />

, 1 ≤ m ≤ n − 1.<br />

Thus, we obtain 1 ≤ Λn,m < π2<br />

4 , 0 ≤ m ≤ n.<br />

We apply the Richardson method to the system in (8.3.7) w<strong>it</strong>h θ := 2/(π2 /4 + 1).<br />

This choice minimizes the spectral radius of M = I − θR −1 D, which takes the value<br />

ρ = (π 2 /4 − 1)/(π 2 /4 + 1) ≈ 0.42 . Starting from the in<strong>it</strong>ial guess ¯p (0) = R −1 ¯q<br />

(see (7.6.2)), one achieves the exact solution to the system (to machine accuracy in<br />

double precision) in about twenty <strong>it</strong>erations. Though we do not have access to the<br />

explic<strong>it</strong> expression of the precond<strong>it</strong>ioned eigenvalues for different values of α and β,<br />

the behavior is more or less the same.<br />

We note that, in practice, the matrix R −1 D is not actually required. Each <strong>it</strong>eration<br />

consists of two steps. In the first step, we evaluate the matrix-vector multiplication<br />

¯p → D¯p. This can be carried out w<strong>it</strong>h the help of the FFT (see section 4.3) in the<br />

Chebyshev case. In the second step, we compute D¯p → R −1 D¯p by solving a tridiagonal<br />

linear system. In this way, we avoid storing the whole matrix R −1 . Moreover, the cost<br />

of this computation is proportional to n (see isaacson and keller(1966), p.55).<br />

Symmetric precond<strong>it</strong>ioners, based on fin<strong>it</strong>e element discretizations, have been pro-<br />

posed in deville and mund(1985) and in canuto and quarteroni(1985). They allow<br />

us to apply a variation of the conjugate gradient <strong>it</strong>erative method for solving (8.3.7).<br />

A standard trick is to accelerate the convergence by updating the parameter θ at<br />

each <strong>it</strong>eration. This can be done in several ways. For brev<strong>it</strong>y, we do not investigate this<br />

here. A survey of the most used precond<strong>it</strong>ioned <strong>it</strong>erative techniques in spectral methods<br />

is given in canuto, hussaini, quarteroni and zang(1988), p.148. Many practical<br />

suggestions are provided in boyd(1989).<br />

We note that R is ill-cond<strong>it</strong>ioned. This introduces small rounding errors when<br />

evaluating R −1 . Hence, after the precond<strong>it</strong>ioned Richardson method has reached a<br />

steady solution, we suggest that <strong>it</strong> be refined by performing a few <strong>it</strong>erations of the<br />

unprecond<strong>it</strong>ioned Richardson method to remove the rounding errors.


Eigenvalue Analysis 171<br />

The matrix in (7.4.11) is obtained after eliminating the first and the last unknowns<br />

from system (7.4.10). A (n − 1) × (n − 1) precond<strong>it</strong>ioner for this matrix is obtained<br />

by deleting the first and the last columns and rows from the matrix in (8.4.1). The<br />

precond<strong>it</strong>ioned eigenvalues Λn,m, 1 ≤ m ≤ n−1, coincide w<strong>it</strong>h those considered above.<br />

The same precond<strong>it</strong>ioner R in (8.4.1) can be used for the matrix of system (7.4.12).<br />

The resulting precond<strong>it</strong>ioned spectrum does not seem to be much affected by lower-<br />

order terms, provided the coefficients A and B are not too large. As an alternative, one<br />

can use a precond<strong>it</strong>ioner based on a centered fin<strong>it</strong>e-difference discretization of the whole<br />

differential operator − d2<br />

dx2 + A d<br />

dx + B (see orszag(1980)).<br />

For other types of boundary cond<strong>it</strong>ions we could run into some difficulty. Let<br />

D denote the matrix of system (7.4.13). In this s<strong>it</strong>uation, tridiagonal fin<strong>it</strong>e-difference<br />

precond<strong>it</strong>ioners seem less effective. All the precond<strong>it</strong>ioned eigenvalues satisfy relation<br />

(8.4.2), except for one of them, which is real, pos<strong>it</strong>ive, but tends to zero as 1/n 2 . This<br />

badly affects the cond<strong>it</strong>ion number. A more appropriate (n+1)×(n+1) precond<strong>it</strong>ioner<br />

R := {rij} 0≤i≤n<br />

0≤j≤n<br />

⎪⎨<br />

(8.4.4) rij :=<br />

is obtained as follows:<br />

⎧<br />

−d (1)<br />

ij<br />

−1<br />

h (n)<br />

i ˆ h (n)<br />

i<br />

2<br />

h (n)<br />

i+1h(n) i<br />

−1<br />

h (n)<br />

i+1 ˆ h (n)<br />

i<br />

d (1)<br />

ij<br />

if i = 0, 0 ≤ j ≤ n,<br />

⎪⎩<br />

0 elsewhere.<br />

if 1 ≤ i = j + 1 ≤ n − 1,<br />

+ µ if 1 ≤ i = j ≤ n − 1,<br />

if 1 ≤ i = j − 1 ≤ n − 1,<br />

if i = n, 0 ≤ j ≤ n,<br />

All the resulting precond<strong>it</strong>ioned eigenvalues are now well-behaved. Unfortunately, the<br />

matrix R is sparse but not banded. However, <strong>it</strong> can be decomposed by a product of


172 Polynomial Approximation of Differential Equations<br />

three well-structured matrices, so that the global cost of the computation ¯p → R −1 ¯p<br />

is proportional to n (see bertoluzza and funaro(1991)). Slight modifications to<br />

(8.4.4) lead to a su<strong>it</strong>able precond<strong>it</strong>ioner for the matrix of system (7.4.16).<br />

Fin<strong>it</strong>e-difference precond<strong>it</strong>ioning matrices for fourth-order problems have been in-<br />

vestigated in funaro and heinrichs(1990), both for the boundary-value problems<br />

(7.4.19) and (7.4.23). The analysis has been carried out in the Chebyshev case, but<br />

similar conclusions also apply to other s<strong>it</strong>uations. A l<strong>it</strong>tle care has to be payed when<br />

dealing w<strong>it</strong>h the fin<strong>it</strong>e-difference schemes at the boundary points. Before precond<strong>it</strong>ion-<br />

ing, the ratio between the largest and the smallest eigenvalues is proportional to n 8 .<br />

The precond<strong>it</strong>ioned eigenvalues satisfy relation (8.4.2). In particular, in the Chebyshev<br />

case, we can take c = π4<br />

16 ≈ 6.08 .<br />

For matrices related to problems defined in unbounded domains, we can use the<br />

same arguments. However, since <strong>it</strong> is not advisable to work w<strong>it</strong>h high degree Laguerre<br />

or Herm<strong>it</strong>e polynomials (see sections 1.6 and 1.7), the matrices are not very large in<br />

practical applications. Therefore, we suggest direct methods to solve these systems.<br />

8.5 Precond<strong>it</strong>ioners for first-order operators<br />

Denote by D the (n + 1) × (n + 1) matrix corresponding to the system (7.4.1). Here,<br />

the idea of using fin<strong>it</strong>e-differences to construct a precond<strong>it</strong>ioner for D does not offer<br />

the same features emphasized in section 8.4. For instance, let R := {rij} 0≤i≤n<br />

0≤j≤n<br />

defined as follows:<br />

−1<br />

⎪⎨<br />

(8.5.1) rij :=<br />

⎧<br />

1 if i = j = 0 ,<br />

h (n)<br />

i<br />

if 1 ≤ i = j + 1 ≤ n,<br />

1<br />

⎪⎩<br />

h (n)<br />

if 1 ≤ i = j ≤ n,<br />

i<br />

0 elsewhere.<br />

be


Eigenvalue Analysis 173<br />

Now, we have a bidiagonal precond<strong>it</strong>ioner. Though the spread of eigenvalues of the<br />

precond<strong>it</strong>ioned matrix is reduced, we are still far from the results of the previous section.<br />

The eigenvalues of R −1 D are complex w<strong>it</strong>h pos<strong>it</strong>ive real part, but they approach the<br />

imaginary axis w<strong>it</strong>h increasing n.<br />

A more effective precond<strong>it</strong>ioner was proposed in funaro(1987). First introduce the<br />

linear operator T : R n → R n . For any polynomial p ∈ Pn−1, the operator T maps<br />

the vector {p(ξ (n)<br />

j )}1≤j≤n into the vector {p(η (n)<br />

i )}1≤i≤n. The way to perform this<br />

transformation is described by formula (4.4.1). The coefficients l (n)<br />

j (η (n)<br />

i ), 1 ≤ i ≤ n,<br />

1 ≤ j ≤ n, are the entries of the n × n matrix relative to the mapping T. Then, we<br />

define Z := {zij} 0≤i≤n , to be the (n + 1) × (n + 1) matrix<br />

0≤j≤n<br />

⎧<br />

1<br />

⎪⎨<br />

if i = j = 0,<br />

(8.5.2) zij :=<br />

⎪⎩<br />

0 elsewhere.<br />

l (n)<br />

j (η (n)<br />

i ) if 1 ≤ i ≤ n, 1 ≤ j ≤ n,<br />

We claim that ZR is a good precond<strong>it</strong>ioning matrix for D. We remark that Z is a<br />

full matrix. On the other hand, we can find explic<strong>it</strong>ly the entries of Z −1 . Indeed, one<br />

easy checks the relation<br />

(8.5.3) p(ξ (n)<br />

i ) =<br />

n<br />

j=1<br />

which is true for any p ∈ Pn−1.<br />

p(η (n)<br />

j )<br />

1 + η(n)<br />

j<br />

1 + ξ (n)<br />

i<br />

˜ l (n)<br />

j (ξ (n)<br />

i ), 1 ≤ i ≤ n,<br />

Clearly, (8.5.3) is the inverse of the transformation T. Noting that (ZR) −1 = R −1 Z −1 ,<br />

we can evaluate ¯p → (ZR) −1 ¯p w<strong>it</strong>h a cost proportional to n 2 . The computing time<br />

is further reduced, in the Chebyshev case, by using the FFT (see section 4.4). Let us<br />

examine the eigenvalues Λn,m, 0 ≤ m ≤ n, of (ZR) −1 D. We have a precise answer in<br />

the Chebyshev case. Actually, one has<br />

(8.5.4) Λn,0 = 1, Λn,m = m<br />

Therefore, we get 1 ≤ Λn,m < π<br />

2 , 0 ≤ m ≤ n.<br />

π sin 2n<br />

sin mπ , 1 ≤ m ≤ n.<br />

2n


174 Polynomial Approximation of Differential Equations<br />

The improvement is explained by noting that the fin<strong>it</strong>e-differences operator R is a<br />

good approximation of the operator Z−1D, which maps the values p(η (n)<br />

j ), 0 ≤ j ≤ n,<br />

into the values p ′ (ξ (n)<br />

i ) ≈ (p(η (n)<br />

i ) − p(η (n)<br />

i−1 ))/h(n) i , 1 ≤ i ≤ n. Thus, R−1 (Z−1D) is<br />

very close to the ident<strong>it</strong>y operator. For different values of α and β, the results are<br />

similar.<br />

W<strong>it</strong>h minor modifications, the same precond<strong>it</strong>ioner can be used to deal w<strong>it</strong>h the<br />

matrix of system (7.4.4). For more details, we refer the reader to funaro and got-<br />

tlieb(1988).<br />

Though the results obtained w<strong>it</strong>h the precond<strong>it</strong>ioning matrix ZR are impressive,<br />

for more general problems, such as (7.4.7), the use of a full precond<strong>it</strong>ioner is not rec-<br />

ommended. For example, consider A ≡ 1 in (7.4.7). The matrix ZR + I turns<br />

out to be an appropriate precond<strong>it</strong>ioner for D + I, i.e. κ((ZR + I) −1 (D + I)) is<br />

close to 1. Nevertheless, we cannot cheaply invert ZR + I. An alternative is to ap-<br />

proximate the operator Z by a banded matrix ˆ Z. In this way ˆ ZR + I will be<br />

also banded. This idea has been developed in funaro and rothman(1989). In that<br />

paper, the operator T is subst<strong>it</strong>uted by a mapping ˆ T, which extrapolates the point<br />

values p(η (n)<br />

i ), 1 ≤ i ≤ n, w<strong>it</strong>h the help of second-degree polynomials. The resulting<br />

precond<strong>it</strong>ioner is four-diagonal and the eigenvalues of the matrix ( ˆ ZR + I) −1 (D + I)<br />

are complex, but clustered in a neighborhood of the real un<strong>it</strong>y. The precond<strong>it</strong>ioned<br />

cond<strong>it</strong>ion number seems to be su<strong>it</strong>able for numerical applications.<br />

8.6 Convergence of the eigenvalues<br />

We are now ready to consider the first application of spectral methods to solve several<br />

problems in physics. For example, let us analyze the following eigenvalue problem:<br />

(8.6.1)<br />

⎧<br />

⎨ −φ ′′ = λ φ in I =] − 1,1[,<br />

⎩<br />

φ(−1) = φ(1) = 0.


Eigenvalue Analysis 175<br />

It is well-known that (8.6.1) has a countable set of pos<strong>it</strong>ive eigenvalues λm = π2<br />

4 m2 ,<br />

m ≥ 1, corresponding to the normalized eigenfunctions: φm(x) = sin π<br />

2<br />

m even, φm(x) = cos π<br />

2<br />

mx, x ∈ Ī,<br />

mx, x ∈ Ī, m odd. Besides, for any n ≥ 2, we have the<br />

eigenvalues λn,m, 1 ≤ m ≤ n −1, of problem (8.2.1). These will be assumed to be real,<br />

distinct and in increasing order.<br />

We are concerned w<strong>it</strong>h the asymptotic behavior of the λn,m’s when n tends to<br />

infin<strong>it</strong>y. Actually, we expect the following convergence result:<br />

(8.6.2) ∀m ≥ 1, lim<br />

n→+∞ λn,m = λm.<br />

This would suggest a technique for approximating the eigenvalues of differential opera-<br />

tors, which is a crucial problem in many applications of mathematical physics (see for<br />

instance courant and hilbert(1953), Vol.1). This procedure is investigated for the<br />

solution of a large number of problems in calogero and franco (1985). In that pa-<br />

per, discretizations of the derivative operators (denoted by Lagrangian differentiations)<br />

are obtained starting from an arb<strong>it</strong>rary set of nodes, not necessarily related to Gauss<br />

type formulas. Heuristic estimates of the rate of convergence of the lim<strong>it</strong> in (8.6.2) are<br />

presented in calogero(1983) for equispaced nodes and durand(1985) for Gauss type<br />

nodes. According to these papers the convergence is exponential, and, as pointed out by<br />

the authors in their numerical experiments, the results are defin<strong>it</strong>ely better than those<br />

obtained by other techniques, such as the fin<strong>it</strong>e-differences method.<br />

Remarks about the behavior of the eigenvalues in the Chebyshev case can be found<br />

in weideman and trefethen(1988). The authors make the following observation.<br />

For any fixed n ≥ 2, about two thirds of the eigenvalues in (8.2.1) are very close to<br />

the corresponding eigenvalues in (8.6.1). The remaining part of the spectrum deviates<br />

sharply. This behavior justifies the estimate (8.3.6), since λn,n−1 turns out to be<br />

proportional to n 4 , while λn just grows like n 2 . This fact is explained by the poor<br />

approximation properties of the eigenfunctions w<strong>it</strong>h high frequency oscillations, where<br />

the number of nodes is not sufficient to recover a satisfactory resolution (see section<br />

6.8). This does not imply that (8.6.2) is false. In fact, for fixed m, the eigenvalue<br />

λn,m begins to converge for sufficiently large n.


176 Polynomial Approximation of Differential Equations<br />

In figure 8.6.1, for the Legendre case, in the interval [0,600], we show the exact eigen-<br />

values λm, 1 ≤ m ≤ 15 (first line) and the computed eigenvalues λn,m, 1 ≤ m ≤ n−1<br />

(next lines), when n varies from 6 to 11.<br />

Figure 8.6.1 - Behavior of the eigenvalues of problem<br />

(8.2.1) in the Legendre case when 6 ≤ n ≤ 11.<br />

In order to check (8.3.6), we first need the following result. We recall that the<br />

polynomial space P 0 n, n ≥ 2, is defined in (6.4.1).<br />

Lemma 8.6.1 - Let −1 < α < 1 and −1 < β < 1. Then we can find two constants<br />

C1 > 0, C2 > 0, such that, for any n ≥ 2 and p ∈ P 0 n, one has<br />

(8.6.3)<br />

<br />

p<br />

<br />

1 − x2 <br />

<br />

w<br />

≤ C1 p ′ w,


Eigenvalue Analysis 177<br />

(8.6.4)<br />

1<br />

−1<br />

p ′ (pw) ′ <br />

dx<br />

1<br />

2<br />

≤ C2 p ′ w.<br />

Proof - Inequal<strong>it</strong>y (8.6.3) is a byproduct of (5.7.5). We provide however some details.<br />

W<strong>it</strong>h the help of the Schwarz inequal<strong>it</strong>y, one obtains<br />

(8.6.5)<br />

≤<br />

<br />

p<br />

<br />

1 − x2 <br />

<br />

2<br />

w<br />

=<br />

1<br />

−1<br />

x<br />

−1<br />

1 x<br />

[p<br />

−1 −1<br />

′ (s)] 2 x<br />

w(s)ds ·<br />

−1<br />

≤ p ′ 2 1 x<br />

w<br />

−1 −1<br />

p ′ 2 w(x)<br />

(s)ds<br />

(1 − x2 dx<br />

) 2<br />

1<br />

w(s) ds<br />

1<br />

w(s) ds<br />

<br />

w(x)<br />

w(x)<br />

(1 − x2 dx<br />

) 2<br />

(1 − x2 dx.<br />

) 2<br />

The last integral in (8.6.5) is fin<strong>it</strong>e, by virtue of the hypotheses on α and β.<br />

Concerning (8.6.4), the Schwarz inequal<strong>it</strong>y yields<br />

(8.6.6)<br />

1<br />

p<br />

−1<br />

′ (pw) ′ dx = p ′ 2 w +<br />

≤ p ′ 2 w +<br />

1<br />

−1<br />

1<br />

−1<br />

<br />

p w′<br />

<br />

√ (p<br />

w<br />

′√ w) dx<br />

p 2 (w ′ ) 2 w −1 <br />

dx<br />

1<br />

2<br />

p ′ w.<br />

By noting that (w ′ ) 2 w −1 ≤ 4(|α| + |β|) 2 (1 − x 2 ) −2 w, we can use (8.6.3) to conclude.<br />

Inequal<strong>it</strong>y (8.3.6) is now straightforward. Starting from relation (8.2.10), we use lemma<br />

8.2.1, inequal<strong>it</strong>y (8.6.3) and theorem 3.8.2, to prove that |λn,m| ≥ c1. On the contrary,<br />

from (8.6.4), (6.3.6) and theorem 3.8.2, we obtain |λn,m| ≤ c2n 4 .<br />

As far as we know, a full analysis on the convergence of both the eigenvalues and the<br />

relative eigenfunctions is not yet available. A general approach, based on the properties<br />

of compact operators, is studied in vainikko(1964), vainikko(1967) and osborn(1975),<br />

as well as in many other papers. A self-contained expos<strong>it</strong>ion w<strong>it</strong>h a comprehensive list


178 Polynomial Approximation of Differential Equations<br />

of references is provided in chatelin(1983), where hints for the treatment of problems<br />

similar to the one considered here are given in chapter 4B.<br />

We can say something more about the Legendre case. Combining the proof of theo-<br />

rem 2.2.1 w<strong>it</strong>h that of theorem 8.2.2, one checks that, for α = β = 0, the eigenfunctions<br />

pn,m, 1 ≤ m ≤ n − 1, as well as their derivatives, are orthogonal w<strong>it</strong>h respect to the<br />

inner product (·, ·)w,n (see section 3.8). This is true if we assume that the eigenvalues<br />

λn,m, 1 ≤ m ≤ n − 1, are distinct.<br />

Let Xm, 1 ≤ m ≤ n − 1, denote a subspace of P 0 n of dimension m. Then, by<br />

virtue of the orthogonal<strong>it</strong>y of the eigenfunctions, the following characterization, known<br />

as Min-Max principle, holds:<br />

(8.6.7) λn,m = min<br />

Xm⊂P 0 n<br />

<br />

max<br />

p∈Xm<br />

p≡0<br />

p ′ 2 w<br />

p 2 w,n<br />

<br />

, 1 ≤ m ≤ n − 1.<br />

Applications of formula (8.6.7), in the field of fin<strong>it</strong>e-element approximations, are pre-<br />

sented for instance in strang and fix(1973), chapter six.<br />

Other s<strong>it</strong>uations may be considered. The problem<br />

⎧<br />

⎨ −φ<br />

(8.6.8)<br />

⎩<br />

′′ + µφ = λ φ in I =] − 1,1[, µ > 0,<br />

φ ′ (−1) = φ ′ (1) = 0,<br />

has a countable set of eigenvalues λm := π2<br />

4 m2 + µ, m ∈ N. These are effectively<br />

approximated by the eigenvalues of problem (8.2.13). We only remark that two of the<br />

λn,m’s are proportional to the parameter γ and are out of interest.<br />

Different arguments apply to problem (8.2.12). Here, the λn,m’s do not approach<br />

the eigenvalues of problem (8.6.8). If we replace the second equation in (8.2.12) by<br />

p ′ n,m(η (n)<br />

0 ) = λn,mpn,m(η (n)<br />

0 ), we get a negative eigenvalue, say λn,0. Another eigen-<br />

value, say λn,1, is such that limn→+∞(λn,0 + λn,1) = −1. The remaining eigenvalues<br />

approximate those of problem (8.6.8), w<strong>it</strong>h the exception of λ0 = µ.<br />

These considerations demonstrate the importance of the treatment of boundary<br />

cond<strong>it</strong>ions in spectral methods. A non correct specification may give raise to unexpected<br />

results. Though this is true in all numerical methods, the s<strong>it</strong>uation seems more crucial


Eigenvalue Analysis 179<br />

for spectral methods, due to their global nature. On the other hand, if the boundary<br />

cond<strong>it</strong>ions are well treated, results are really compet<strong>it</strong>ive. An example of a second-order<br />

eigenvalue problem is examined in section 12.4.<br />

Let us finally consider the eigenvalues associated to the system (7.4.22). These<br />

converge to the eigenvalues of the problem<br />

⎧<br />

⎨φ<br />

(8.6.9)<br />

⎩<br />

IV = λ φ in I =] − 1,1[,<br />

φ(±1) = φ ′ (±1) = 0.<br />

Here, λ assumes a countable number of pos<strong>it</strong>ive values corresponding to the roots of<br />

the equation cosh2 4√ λ cos2 4√ λ = 1 (see courant and hilbert(1953), Vol.1, p.296).<br />

No asymptotic estimates can be recovered for the eigenvalues of problem (8.1.1).<br />

In fact, the equation φ ′ = λφ in ] − 1,1[, φ(−1) = σ, has only the solution φ ≡ 0<br />

when σ = 0.<br />

8.7 Multigrid method<br />

We already noticed that the speed of convergence of the Richardson method strongly<br />

depends on the location of the eigenvalues λm, 1 ≤ m ≤ n, of the matrix D in (7.6.1).<br />

Let ¯pm ∈ C n , 1 ≤ m ≤ n, denote the eigenvectors of D, su<strong>it</strong>ably normalized. In the<br />

computation ¯p → M ¯p, different choices of the parameter θ will result in a different<br />

treatment of the components (modes) of the vector ¯p, along the directions selected<br />

by ¯pm, 1 ≤ m ≤ n. For instance, let us assume that the λm’s are real, distinct,<br />

pos<strong>it</strong>ive and in increasing order. By examining relation (8.3.5), we deduce that the<br />

maximum damping of the components corresponding to a large integer m (high modes),<br />

is obtained for small values of θ. Conversely, low modes decay fast for larger values<br />

of the parameter θ, provided the assumptions in (7.6.3) are satisfied. Therefore, for<br />

ill-cond<strong>it</strong>ioned matrices we cannot expect a strong reduction of all the modes.<br />

This can be partly overcome by using the multigrid method. We illustrate this procedure<br />

w<strong>it</strong>h an example. We denote by D (n) , n ≥ 2, the (n+1)×(n+1) matrix of the system


180 Polynomial Approximation of Differential Equations<br />

in (7.4.9). This is obtained starting from the set of grid-points η (n)<br />

j , 0 ≤ j ≤ n (finegrid).<br />

Another set of points in [−1,1] (coarse-grid) is given by the nodes η (k)<br />

j , 0 ≤ j ≤ k,<br />

where 2 ≤ k ≤ n. A polynomial p ∈ Pn, determined by <strong>it</strong>s values on the fine-grid,<br />

is evaluated at the coarse-grid by using formula (3.2.7). On the other hand, we can go<br />

from the coarse-grid to the fine-grid by extrapolation, i.e.<br />

k<br />

(8.7.1) p(η (n)<br />

i ) =<br />

j=0<br />

p(η (k)<br />

j ) ˜ l (k)<br />

j (η (n)<br />

i ), 0 ≤ i ≤ n.<br />

In short, the philosophy of the multigrid method is the following. Since the eigenvalues<br />

of D (n) are not available (the cost of their computation is very high), an alternative<br />

to varying the parameter θ is to accelerate the convergence by using different grids.<br />

A good treatment of low modes is obtained by applying the Richardson method to<br />

the matrix D (k) (k small) defined on the coarse-grid. Then, we can sw<strong>it</strong>ch to the<br />

fine-grid by (8.7.1), and work w<strong>it</strong>h the matrix D (n) to damp the high modes w<strong>it</strong>h<br />

a relatively small θ. In general, the algor<strong>it</strong>hm applies to different sets of grid-points,<br />

selected in turn, according to some prescribed rule (the most classical one is the V-<br />

cycle). There are many ways to advance the method. These depend on the problem<br />

and the competence of the user. A comparative analysis is given in heinrichs(1988).<br />

Other results are provided in muñoz(1990). Early applications w<strong>it</strong>hin the framework of<br />

spectral methods, for more specialized problems, are examined, for instance, in zang,<br />

wong and hussaini(1982), streett, zang and hussaini(1985), brandt, fulton<br />

and taylor(1985), zang and hussaini(1986). For a general overview of the method,<br />

we refer to hackbusch(1985).<br />

Using the multigrid method for the precond<strong>it</strong>ioned systems in place of a plain<br />

precond<strong>it</strong>ioned <strong>it</strong>erative scheme, does not improve very much the rate of convergence<br />

for problems in one dimension. More interesting are the applications in two or three<br />

dimensions. Performances also depend on the veloc<strong>it</strong>y and the cost to transfer the data<br />

from one grid to the next. Particular efficiency is obtained in the Chebyshev case. When<br />

n is even and k = n/2, recalling relation (3.8.16), all the points of the coarse-grid are<br />

also points of the fine-grid. Furthermore, the inverse formula (8.7.1) is implemented via<br />

FFT.


9<br />

ORDINARY<br />

DIFFERENTIAL EQUATIONS<br />

The large amount of material accumulated in the previous chapters is now directed<br />

towards the study of the convergence of approximate solutions to differential equations.<br />

Beginning w<strong>it</strong>h the simplest linear equations, we then extend our analysis to other, more<br />

complex, problems. Throughout this chapter, however, we remain focused on different<br />

techniques, rather than on solving complicated equations.<br />

9.1 General considerations<br />

Assume that I ⊂ R is a bounded open interval. W<strong>it</strong>hout loss of general<strong>it</strong>y we set<br />

I :=] − 1,1[. In fact, any other interval is mapped into ] − 1,1[ by a linear change of<br />

variables and a translation. Note that the space of polynomials Pn, n ∈ N, is invariant<br />

under these transformations.<br />

Let f : Ī → R be a given continuous function. A very simple differential equation is<br />

(9.1.1)<br />

⎧<br />

⎨ U ′ = f in ] − 1,1],<br />

⎩<br />

U(−1) = σ,


182 Polynomial Approximation of Differential Equations<br />

where σ ∈ R , and U :<br />

explic<strong>it</strong>ly, i.e.<br />

(9.1.2) U(x) = σ +<br />

Ī → R is the unknown. Of course, we know the solution<br />

x<br />

−1<br />

f(s) ds, ∀x ∈ Ī.<br />

We are concerned w<strong>it</strong>h the convergence analysis of polynomial approximations of U.<br />

We give an overview of different techniques in the coming sections.<br />

A generalization of (9.1.1) is<br />

(9.1.3)<br />

⎧<br />

⎨U<br />

′ + AU = f in ] − 1,1],<br />

⎩<br />

U(−1) = σ,<br />

where A : Ī → R is a given continuous function.<br />

Other elementary problems can be studied. We are mainly interested in boundary-<br />

value problems. Two typical examples of second-order equations are<br />

(9.1.4)<br />

(9.1.5)<br />

⎧<br />

⎨ −U ′′ = f in I,<br />

⎩<br />

U(−1) = σ1, U(1) = σ2,<br />

⎧<br />

⎨ −U ′′ + µU = f in I,<br />

⎩<br />

U ′ (−1) = σ1, U ′ (1) = σ2,<br />

where σ1, σ2 ∈ R and µ > 0. In particular, (9.1.4) and (9.1.5) are known as Dirichlet<br />

problem and Neumann problem respectively.<br />

W<strong>it</strong>hin the framework of fourth-order problems, we consider for instance the equation<br />

(9.1.6)<br />

w<strong>it</strong>h σi ∈ R, 1 ≤ i ≤ 4.<br />

⎧<br />

U<br />

⎪⎨<br />

⎪⎩<br />

IV = f in I,<br />

U(−1) = σ1, U(1) = σ2,<br />

U ′ (−1) = σ3, U ′ (1) = σ4,


Ordinary Differential Equations 183<br />

Theoretical considerations of the problems proposed above are presented in many<br />

books on ordinary differential equations. We refer for instance to golomb and shanks<br />

(1965), brauer and nohel(1967), simmons(1972), etc.. Once we gain enough experi-<br />

ence w<strong>it</strong>h the numerical treatment of these elementary equations, we will turn to more<br />

realistic problems. Several non trivial examples are discussed in chapter twelve.<br />

9.2 Approximation of linear equations<br />

A technique to approximate the solution of problem (9.1.1) is the so called tau method.<br />

Early applications were considered in lanczos(1956). Theoretical developments are<br />

given for instance in gottlieb and orszag(1977), for Chebyshev and Legendre ex-<br />

pansions. The unknown function U is approximated by a sequence of polynomials<br />

pn ∈ Pn, n ≥ 1, which are expressed in the frequency space relative to some Jacobi<br />

weight function w. For any n ≥ 1, these are required to be solutions of the following<br />

problem:<br />

(9.2.1)<br />

⎧<br />

⎨<br />

p ′ n = Πw,n−1f in ] − 1,1],<br />

⎩<br />

pn(−1) = σ.<br />

The operator Πw,n, n ∈ N, is the same operator introduced in section 2.4. In this<br />

scheme the boundary cond<strong>it</strong>ion is the same as in (9.1.1), but the function f is replaced<br />

by a truncation of the series (6.2.11). In practice, we project the differential equation<br />

in the space Pn−1. Following the suggestions provided in section 7.3, where now<br />

q := Πw,n−1f, we can reduce (9.2.1) to a linear system. The vector on the right-hand<br />

side contains the first n Fourier coefficients of the function f and the datum σ. The<br />

unknown vector consists of the Fourier coefficients of the polynomial pn. Note that<br />

these do not coincide w<strong>it</strong>h the first n + 1 Fourier coefficients of the function U.<br />

The question is to check whether limn→+∞ pn = U, and how to interpret this lim<strong>it</strong>.<br />

By (9.1.2), we easily get


184 Polynomial Approximation of Differential Equations<br />

(9.2.2) U(x) − pn(x) =<br />

x<br />

−1<br />

(f − Πw,n−1f)(s) ds, ∀x ∈ Ī.<br />

At this point, an estimate of the rate of convergence to zero of the error U −pn, n ≥ 1,<br />

in some su<strong>it</strong>able norm, is straightforward after recalling the results of section 6.2. For<br />

example, when −1 < α < 1 and −1 < β < 1, <strong>it</strong> is easy to obtain uniform convergence<br />

w<strong>it</strong>h the help of the Schwarz inequal<strong>it</strong>y and of theorem 6.2.4. Indeed, for k ∈ N, one<br />

has<br />

(9.2.3) sup<br />

x∈[−1,1]<br />

|U(x) − pn(x)| ≤ C<br />

k 1<br />

fH k<br />

w n<br />

(I), ∀n > k.<br />

The space H k w(I), k ∈ N, is defined in (5.7.1). As already observed in chapter six, the<br />

more regular the function f, the faster the decay of the error is.<br />

Another approach is to construct the approximation scheme in the physical space.<br />

This leads to the collocation (or pseudospectral) method. Now, the approximating poly-<br />

nomials pn ∈ Pn, n ≥ 1, satisfy<br />

⎧<br />

⎨<br />

(9.2.4)<br />

⎩<br />

p ′ n(η (n)<br />

i ) = f(η (n)<br />

i ) 1 ≤ i ≤ n,<br />

pn(η (n)<br />

0 ) = σ.<br />

Here, we impose the boundary cond<strong>it</strong>ion at x = −1 and we collocate the differential<br />

equation at the nodes η (n)<br />

i , 1 ≤ i ≤ n. We are left w<strong>it</strong>h a linear system in the unknowns<br />

pn(η (n)<br />

j ), 0 ≤ j ≤ n. The corresponding matrix is obtained following the guideline of<br />

section 7.4. We remark that the computed values pn(η (n)<br />

j ), 0 ≤ j ≤ n, do not coincide<br />

in general w<strong>it</strong>h the values U(η (n)<br />

j ), 0 ≤ j ≤ n.<br />

We can study the convergence of the collocation method by arguing as follows. We<br />

define r(x) := 1 + x, x ∈ Ī. Then, we note that (9.2.4) is equivalent to<br />

(9.2.5)<br />

⎧<br />

⎨p<br />

⎩<br />

′ n = [ Ĩw,n(fr)]r −1 pn(−1) = σ,<br />

in ] − 1,1],


Ordinary Differential Equations 185<br />

where Ĩw,n, n ≥ 1, is the interpolation operator at the points η (n)<br />

j , 0 ≤ j ≤ n (see<br />

section 3.3). It turns out that [ Ĩw,n(fr)]r −1 is the interpolant of f at the nodes<br />

η (n)<br />

j , 1 ≤ j ≤ n. This shows the equivalence of (9.2.5) w<strong>it</strong>h (9.2.4). The analysis of the<br />

error U −pn, n ≥ 1, is then recovered from the estimates of the error [fr− Ĩw,n(fr)]r −1 .<br />

These results have been developed in section 6.6.<br />

The two approximation methods described here differ in the treatment of the func-<br />

tion f. Projection and interpolation operators have been used respectively. A small<br />

difference is due to the aliasing error (see section 4.2), but the approximating solutions<br />

behave similarly. The first approach is more effective in terms of computational cost.<br />

Tau and collocation methods are respectively used to approximate the solution of<br />

problem (9.1.3) according to the following schemes:<br />

(9.2.6)<br />

(9.2.7)<br />

⎧<br />

⎨<br />

⎩<br />

⎧<br />

⎨<br />

p ′ n + Πw,n−1(Apn) = Πw,n−1f in ] − 1,1],<br />

⎩<br />

pn(−1) = σ ,<br />

p ′ n(η (n)<br />

i ) + A(η (n)<br />

i )pn(η (n)<br />

i ) = f(η (n)<br />

i ) 1 ≤ i ≤ n,<br />

pn(η (n)<br />

0 ) = σ ,<br />

pn ∈ Pn, n ≥ 1.<br />

The corresponding linear systems are given by (7.3.3) and (7.4.7). Here, the second<br />

approach is preferable, having the coefficients of the matrix relative to the tau method<br />

a complicate expression, for a non-constant function A.<br />

The analysis of convergence is now more involved and general results are not avail-<br />

able. We examine for instance the collocation method in the Legendre case. We recall<br />

that the norm in the space L 2 w(I) is defined in (5.2.4).


186 Polynomial Approximation of Differential Equations<br />

Theorem 9.2.1 - Let α = β = 0 and A : Ī → R, be a continuous pos<strong>it</strong>ive function<br />

satisfying A(x) ≥ ǫ, ∀x ∈ Ī, for a given ǫ > 0. Then<br />

(9.2.8) lim<br />

n→+∞ pn − U L 2 w (I) = 0,<br />

where pn, n ≥ 1, is the solution of (9.2.7) and U is the solution of (9.1.3).<br />

Proof - We first note that, by the triangle inequal<strong>it</strong>y (2.1.5), one has<br />

(9.2.9) pn − U L 2 w (I) ≤ pn − Ĩw,nU L 2 w (I) + Ĩw,nU − U L 2 w (I), n ≥ 1.<br />

The last term of the right-hand side tends to zero in view of theorem 6.6.1. After defining<br />

sn := pn − Ĩw,nU ∈ Pn, n ≥ 1, we require an estimate for the error sn L 2 w (I), n ≥ 1.<br />

Equations (9.1.3) and (9.2.7) yield<br />

(9.2.10)<br />

⎧<br />

⎨<br />

⎩<br />

s ′ n(η (n)<br />

i ) + A(η (n)<br />

i )sn(η (n)<br />

i ) = (U − Ĩw,nU) ′ (η (n)<br />

i ) 1 ≤ i ≤ n,<br />

sn(η (n)<br />

0 ) = 0.<br />

Using the quadrature formula (3.5.1) w<strong>it</strong>h w ≡ 1, we obtain<br />

(9.2.11)<br />

Next, we observe that<br />

=<br />

<br />

s<br />

I<br />

′ nsn dx +<br />

n<br />

i=0<br />

n<br />

i=0<br />

tions on A and the Schwarz inequal<strong>it</strong>y, one gets<br />

<br />

A(η (n)<br />

i )s 2 n(η (n)<br />

i ) ˜w (n)<br />

i<br />

(U − Ĩw,nU) ′ (η (n)<br />

i ) sn(η (n)<br />

i ) ˜w (n)<br />

i , n ≥ 1.<br />

I s′ nsndx = 1<br />

2 s2 n(1) ≥ 0, n ≥ 1. Therefore, from the assump-<br />

(9.2.12) ǫ snw,n ≤ Ĩw,nU ′ − ( Ĩw,nU) ′ w,n, n ≥ 1,<br />

where we note that Ĩw,nU ′ coincides w<strong>it</strong>h U ′ at the nodes. The norm ·w,n, n ≥ 1, is<br />

given in (3.8.2). By virtue of theorem 3.8.2, this norm is equivalent to the L 2 w(I) norm.<br />

Besides, when U ∈ C1 ( Ī), the right-hand side of (9.2.12) tends to zero. Actually, by<br />

(3.8.6), one has


Ordinary Differential Equations 187<br />

(9.2.13) Ĩw,nU ′ − ( Ĩw,nU) ′ w,n ≤ γ −1<br />

1 Ĩw,nU ′ − ( Ĩw,nU) ′ L 2 w (I)<br />

≤ γ −1<br />

1<br />

<br />

Ĩw,nU ′ − U ′ L 2 w (I) + ( Ĩw,nU − U) ′ L 2 w (I)<br />

<br />

, n ≥ 1.<br />

Finally, we use inequal<strong>it</strong>ies like (6.6.1) and (6.6.8) to estimate the last term. This shows<br />

that sn L 2 w (I) tends to zero for n → +∞. Thus, we get (9.2.8).<br />

As seen from the proof given above, the rate of convergence in (9.2.8) is the typical<br />

one of spectral approximations. In particular, for analytic solutions U, the decay of<br />

the error is exponential. The extension of theorem 9.2.1 to other Jacobi weights is not<br />

straightforward. Convergence estimates for the Chebyshev case can be derived from<br />

the results in solomonoff and turkel(1989). The use of collocation nodes relative to<br />

Gauss type formulas is helpful in the theoretical analysis, since <strong>it</strong> allows us to subst<strong>it</strong>ute<br />

summations w<strong>it</strong>h integrals as in (9.2.11). Unfortunately, when the weight function w<br />

is singular at the boundary, the integrals cannot be manipulated further. This is the<br />

major drawback in the analysis of approximations involving Chebyshev polynomials.<br />

For example, the quant<strong>it</strong>y <br />

I φ′ φwdx, φ ∈ Pn, φ(−1) = 0 is not in general pos<strong>it</strong>ive<br />

when w(x) = 1/ √ 1 − x 2 , x ∈ I (see gottlieb and orszag(1977), p.89).<br />

First, we show the results of numerical experiment when U is a very smooth func-<br />

tion. We choose σ = 0 and A(x) := 1+x2 , x ∈ Ī. The datum f in (9.1.3) is such that<br />

U(x) := sin (1 + x), x ∈ Ī. In table 9.2.1, we give the error En := pn − Ĩw,nU L 2 w (I) for<br />

various n, when pn is obtained w<strong>it</strong>h the scheme (9.2.7) relative to the Legendre nodes.<br />

It is evident that the convergence is extremely fast. Note that, from (9.2.9), the error<br />

pn − U L 2 w (I) is bounded by En plus a term which does not depend on the poly-<br />

nomial pn. Therefore, the examination of En gives sufficient information about the<br />

efficiency of the approximation technique. In add<strong>it</strong>ion, according to formula (3.8.10),<br />

the computation of En is obtained from the values of pn at the nodes. These values<br />

are the actual solutions of (9.2.7), after inverting the linear system corresponding to<br />

(7.4.7).


188 Polynomial Approximation of Differential Equations<br />

n En n En<br />

2 .8249 × 10 −1 8 .3399 × 10 −7<br />

3 .1814 × 10 −1 9 .2784 × 10 −8<br />

4 .1232 × 10 −2 10 .8594 × 10 −10<br />

5 .1697 × 10 −3 11 .5817 × 10 −11<br />

6 .8560 × 10 −5 12 .1511 × 10 −12<br />

7 .8874 × 10 −6 13 .8770 × 10 −14<br />

Table 9.2.1 - Errors relative to the Legendre collocation<br />

approximation of problem (9.1.3).<br />

We consider another example. In figure 9.2.1, we plot the approximating polynomials<br />

pn, n = 4, 10, 16, obtained by the collocation method (9.2.7) at the Legendre nodes.<br />

Figure 9.2.1 - Legendre collocation approximations for n = 4, 10, 16.


Ordinary Differential Equations 189<br />

The exact solution (dashed line in figure 9.2.1) is U(x) := √ 1 + x − 1, x ∈ Ī, corre-<br />

sponding to the data: σ = −1, A ≡ 1<br />

2<br />

and f(x) := 2+x<br />

2 √ 1+x<br />

1 − 2 , x ∈] − 1,1]. We<br />

observe that f ∈ C0 ( Ī) and that U has a singular<strong>it</strong>y in the derivative at x = −1.<br />

Nevertheless, convergence is still achieved. The same experiment, using the collocation<br />

method at the Chebyshev nodes, displays a similar behavior.<br />

The numerical schemes introduced here can be also generalized to other type of<br />

equations. For example, tau and collocation methods applied to problem (9.1.4), give<br />

respectively<br />

(9.2.14)<br />

(9.2.15)<br />

⎧<br />

⎨<br />

⎩<br />

⎧<br />

⎨<br />

−p ′′ n = Πw,n−2f in I,<br />

⎩<br />

pn(−1) = σ1, pn(1) = σ2,<br />

−p ′′ n(η (n)<br />

i ) = f(η (n)<br />

i ) 1 ≤ i ≤ n − 1,<br />

pn(η (n)<br />

0 ) = σ1, pn(η (n)<br />

n ) = σ2,<br />

pn ∈ Pn, n ≥ 2.<br />

The systems corresponding to (9.2.14) an (9.2.15) have been investigated in sections<br />

7.3 and 7.4, respectively. Similar considerations hold for the remaining differential<br />

equations proposed in section 9.1. In all cases, we expect convergence of the sequence of<br />

polynomials pn to the exact solution U, as n tends to infin<strong>it</strong>y. To provide convergence<br />

estimates, <strong>it</strong> is useful to restate the original problems in an appropriate way. We study<br />

how to do this in the next sections.<br />

9.3 The weak formulation<br />

An effective approach for the study of solutions of differential equations, and how to<br />

approximate them, is the variational method. Here, the problem is conveniently for-<br />

mulated in the so-called weak form. For a given equation, there are several ways to<br />

proceed.


190 Polynomial Approximation of Differential Equations<br />

Classical formulations are illustrated for instance in lions and magenes(1972). Alter-<br />

native formulations, specifically addressed to applications in the field of spectral meth-<br />

ods, are introduced in canuto and quarteroni(1981), maday and quarteroni(1981),<br />

canuto and quarteroni(1982b). Following these papers, we present several useful<br />

results. Let us begin by examining the Dirichlet problem (9.1.4). W<strong>it</strong>hout loss of<br />

general<strong>it</strong>y, we can assume that σ1 = σ2 = 0 (homogeneous boundary cond<strong>it</strong>ions). Ac-<br />

tually, solutions corresponding to non-homogeneous boundary cond<strong>it</strong>ions are obtained<br />

by adding to U the function r(x) := 1<br />

2 (1 − x)σ1 + 1<br />

2 (1 + x)σ2, x ∈ Ī.<br />

Let X be a su<strong>it</strong>able functional space, for now the space C ∞ 0 (I) (see section 5.4).<br />

Recall that φ ∈ X satisfies φ(±1) = 0. Thus, one obtains the ident<strong>it</strong>y<br />

<br />

(9.3.1)<br />

U ′ (φw) ′ <br />

dx = fφw dx, ∀φ ∈ X.<br />

I<br />

I<br />

The above relation is obtained by multiplying both sides of equation (9.1.4) by φw and<br />

integrating in I. The final form follows after integration by parts using the cond<strong>it</strong>ion<br />

φ(±1) = 0. In this context, φ is called a test function. It is an easy exercise to check<br />

that, due to the freedom to choose the test functions, we start from (9.3.1) and deduce<br />

the original equation (9.1.4). Thus, one concludes that, when U is solution to (9.3.1)<br />

w<strong>it</strong>h U(±1) = 0, then <strong>it</strong> is also solution to (9.1.4) w<strong>it</strong>h σ1 = σ2 = 0.<br />

The advantage of replacing (9.1.4) w<strong>it</strong>h (9.3.1) is that the new formulation can be<br />

su<strong>it</strong>ably generalized by weakening the hypotheses on the data and the solution. To this<br />

end, we note that, according to (9.3.1), U is not required to have a second derivative.<br />

This regular<strong>it</strong>y assumption is subst<strong>it</strong>uted by appropriate integrabil<strong>it</strong>y cond<strong>it</strong>ions on<br />

U ′ and φ ′ . In order to proceed w<strong>it</strong>h our investigation, we need to use the functional<br />

spaces introduced in section 5.7. Henceforth, w will be the Jacobi weight function in<br />

the ultraspherical case, where ν := α = β satisfies −1 < ν < 1. We shall seek the<br />

solution U in the space H 1 0,w(I) (see (5.7.6)), which is also a natural choice for the<br />

space of test functions. Therefore, we set X ≡ H 1 0,w(I). Finally, we take f ∈ L 2 w(I)<br />

(see (5.2.1)). As we will see, these assumptions are sufficient to guarantee the existence<br />

of the integrals in (9.3.1), which are now considered in the Lebesgue sense.<br />

The next step is to insure that, in sp<strong>it</strong>e of such generalizations, problem (9.3.1) is still


Ordinary Differential Equations 191<br />

well-posed. This means that <strong>it</strong> adm<strong>it</strong>s a unique weak solution U ∈ X. We remark that<br />

two solutions of (9.3.1) which differ on a set of measure zero are in the same class of<br />

equivalence in X.<br />

We first introduce some notation. Consider the bilinear form Bw : X ×X → R defined<br />

as follows:<br />

<br />

(9.3.2) Bw(ψ,φ) :=<br />

ψ<br />

I<br />

′ (φw) ′ dx, ∀ψ,φ ∈ X.<br />

The mapping Bw is linear in each argument (this justifies the term bilinear). We also<br />

define the linear operator Fw : X → R according to<br />

<br />

(9.3.3) Fw(φ) :=<br />

I<br />

fφw dx, ∀φ ∈ X.<br />

W<strong>it</strong>h the above defin<strong>it</strong>ions we restate (9.3.1). We are concerned w<strong>it</strong>h finding U ∈ X<br />

such that<br />

(9.3.4) Bw(U,φ) = Fw(φ), ∀φ ∈ X.<br />

The answer to our question relies on the following general result.<br />

Theorem 9.3.1 (Lax & Milgram) - Let X be a Hilbert space. Let B : X × X → R<br />

be a bilinear form and F : X → R be a linear operator. Let us assume that there exist<br />

three pos<strong>it</strong>ive constants C1,C2,C3 such that<br />

(9.3.5) |B(ψ,φ)| ≤ C1 ψX φX, ∀ψ,φ ∈ X,<br />

(9.3.6) B(ψ,ψ) ≥ C2 ψ 2 X, ∀ψ ∈ X,<br />

(9.3.7) |F(φ)| ≤ C3 φX, ∀φ ∈ X.


192 Polynomial Approximation of Differential Equations<br />

Then, there exists a unique solution U ∈ X of problem<br />

(9.3.8) B(U,φ) = F(φ), ∀φ ∈ X.<br />

Moreover, we can find a pos<strong>it</strong>ive constant C4 > 0 such that<br />

(9.3.9) UX ≤ C4 sup<br />

φ∈X<br />

φ≡0<br />

|F(φ)|<br />

.<br />

φX<br />

The Lax-Milgram theorem is a powerful tool for proving existence and uniqueness for a<br />

large class of linear differential problems. Though the proof is not difficult, we om<strong>it</strong> <strong>it</strong><br />

for simplic<strong>it</strong>y. Applications and extensions have been considered in many publications.<br />

We suggest the following authors: lax and milgram(1954), brezis(1983), brezzi and<br />

gilardi(1987). Inequal<strong>it</strong>y (9.3.6) plays a fundamental role in the proof of theorem<br />

9.3.1. It is in general known as a coerciveness cond<strong>it</strong>ion and indicates that (9.3.8) is an<br />

elliptic problem.<br />

The final step is to check wether the Lax-Milgram theorem can be applied to study<br />

the solution of (9.3.4). It is qu<strong>it</strong>e easy to show that X ≡ H1 0,w(I) ⊂ C0 ( Ī) is a Hilbert<br />

space (see section 5.3). In add<strong>it</strong>ion, inequal<strong>it</strong>ies (9.3.5), (9.3.6) and (9.3.7) are directly<br />

obtained from the following result (we recall that the norm in H 1 0,w(I) is given by<br />

(5.7.7)).<br />

Theorem 9.3.2 - Let ν := α = β w<strong>it</strong>h −1 < ν < 1. Then, we can find three pos<strong>it</strong>ive<br />

constants C1,C2,C3 such that<br />

<br />

<br />

(9.3.10) <br />

ψ ′ (φw) ′ <br />

<br />

dx<br />

≤ C1<br />

(9.3.11)<br />

(9.3.12)<br />

I<br />

<br />

<br />

<br />

<br />

I<br />

<br />

[ψ<br />

I<br />

′ ] 2 w dx<br />

ψ<br />

I<br />

′ (ψw) ′ dx ≥ C2<br />

<br />

<br />

fφw dx<br />

≤ C3 f L 2 w (I)<br />

<br />

I<br />

1<br />

2 <br />

I<br />

[φ ′ ] 2 1<br />

w dx<br />

2<br />

, ∀ψ,φ ∈ H 1 0,w(I),<br />

[ψ ′ ] 2 w dx, ∀ψ ∈ H 1 0,w(I),<br />

<br />

[φ<br />

I<br />

′ ] 2 w dx<br />

1<br />

2<br />

, ∀φ ∈ H 1 0,w(I).


Ordinary Differential Equations 193<br />

Proof - We start by considering ψ,φ ∈ P 0 n ⊂ H 1 0,w(I), for some n ≥ 2. Then, the<br />

first inequal<strong>it</strong>y is obtained by applying the Schwarz inequal<strong>it</strong>y (2.1.7) to the integral<br />

<br />

I (ψ′√ w)[(φw) ′ / √ w]dx. Recalling that (w ′ ) 2 /w ≤ w/(1 − x 2 ) 2 , we conclude using<br />

the results of lemma 8.6.1. The second inequal<strong>it</strong>y is proven in lemma 8.2.1. The third<br />

one comes from the Schwarz inequal<strong>it</strong>y and (5.7.5). The proof of (9.3.10), (9.3.11),<br />

(9.3.12) in H 1 0,w(I) requires a l<strong>it</strong>tle care. In short, one approximates φ ∈ H 1 0,w(I)<br />

by a sequence of polynomials φn ∈ P 0 n, n ∈ N, converging to φ in such a way that<br />

limn→+∞ φ − φn H 1 0,w (I) = 0. This also implies: limn→+∞ φn H 1 0,w (I) = φ H 1 0,w (I).<br />

The propos<strong>it</strong>ion follows by a lim<strong>it</strong> process which is standard in Lebesgue integration<br />

theory (see kolmogorov and fomin(1961)). The details are om<strong>it</strong>ted.<br />

Thus, (9.3.4) is a well-posed problem. In the special case when f ∈ C 0 ( Ī) ⊂ L2 w(I),<br />

the solution U of (9.3.4) is the classical strong solution satisfying (9.1.4).<br />

Generalizations can be examined w<strong>it</strong>h l<strong>it</strong>tle modification. For example, let us<br />

consider the boundary-value problem<br />

(9.3.13)<br />

⎧<br />

⎨<br />

−U ′′ + A1U ′ + A2U = f in I,<br />

⎩<br />

U(±1) = 0,<br />

where A1 : Ī → R and A2 : Ī → R are given bounded integrable functions. Here, we<br />

can replace (9.3.2) by the bilinear form<br />

(9.3.14) Bw(ψ,φ) =<br />

<br />

ψ<br />

I<br />

′ (φw) ′ dx +<br />

<br />

I<br />

[A1ψ ′ φ + A2ψφ]w dx, ∀ψ,φ ∈ X.<br />

One checks that, if the inequal<strong>it</strong>y A2w − 1<br />

2 (A1w) ′ ≥ 0 holds in I, then the hypotheses<br />

of theorem 9.3.1 are satisfied. Therefore, the variational formulation of (9.3.13) adm<strong>it</strong>s<br />

a unique weak solution.


194 Polynomial Approximation of Differential Equations<br />

We now propose a weak formulation of the Neumann problem (9.1.5). Let us<br />

first discuss the Legendre case. We set X ≡ H 1 w(I), where w ≡ 1. We introduce<br />

Bw : X × X → R and Fw : X → R such that<br />

<br />

(9.3.15) Bw(ψ,φ) :=<br />

(9.3.16) Fw(ψ) :=<br />

<br />

I<br />

ψ<br />

I<br />

′ φ ′ dx + µ<br />

<br />

I<br />

ψφ dx, ∀ψ,φ ∈ X,<br />

fφ dx + σ2φ(1) − σ1φ(−1), ∀φ ∈ X.<br />

According to these new defin<strong>it</strong>ions, we consider the problem of finding U ∈ X which<br />

satisfies (9.3.4).<br />

Theorem 9.3.3 - Let w ≡ 1 and X ≡ H 1 w(I). Let Bw and Fw be respectively defined<br />

by (9.3.15) and (9.3.16). Then, problem (9.3.4) has a unique weak solution U ∈ X.<br />

Moreover, if U ∈ C2 ( Ī) , then U is solution of (9.1.5).<br />

Proof - It is an easy matter to verify the hypotheses of theorem 9.3.1. In particular,<br />

to prove (9.3.7) we note that, by (5.7.4), one has |φ(±1)| ≤ φ C 0 (Ī) ≤ K2K −1<br />

1 φX.<br />

Then, we have existence and uniqueness of a weak solution U ∈ X. When U is more<br />

regular, we can wr<strong>it</strong>e<br />

(9.3.17) U ′ (1)φ(1) − U ′ (−1)φ(−1) +<br />

=<br />

<br />

I<br />

<br />

I<br />

[−U ′′ + µU]φ dx<br />

fφ dx + σ2φ(1) − σ1φ(−1), ∀φ ∈ X,<br />

which leads to −U ′′ + µU = f in I, by testing on the functions φ ∈ H 1 0,w(I) ⊂ X.<br />

We can now remove the integrals in (9.3.17) and recover the boundary cond<strong>it</strong>ions<br />

U ′ (−1) = σ1, U ′ (1) = σ2.<br />

We can also study variational formulations of the Neumann problem based on<br />

different Jacobi weight functions. We set X ≡ H 1 w(I), where w is the ultraspherical<br />

weight function w<strong>it</strong>h ν := α = β satisfying −1 < ν < 1. Then, Bw : X × X → R<br />

and Fw : X → R are modified to get


Ordinary Differential Equations 195<br />

(9.3.18) Bw(ψ,φ) :=<br />

(9.3.19) Fw(φ) :=<br />

<br />

<br />

(ψ −<br />

I<br />

˜ ψ) ′ [(φ − ˜ φ)w] ′ dx +<br />

<br />

+ µ ψ(φ −<br />

I<br />

˜ <br />

φ)w dx + µ<br />

I<br />

f(φ − ˜ φ)w dx +<br />

<br />

I<br />

I<br />

<br />

I<br />

˜ψ ′ ˜ φ ′ dx<br />

ψ ˜ φ dx, ∀ψ,φ ∈ X,<br />

f ˜ φ dx + σ2φ(1) − σ1φ(−1), ∀φ ∈ X.<br />

In the above expressions, for a given function χ ∈ X, we have defined ˜χ ∈ X to be<br />

the function ˜χ(x) := 1<br />

1<br />

2 (1 − x)χ(−1) + 2 (1 + x)χ(1), x ∈ Ī. We note that ˜χ ∈ P1<br />

and χ − ˜χ ∈ H 1 0,w(I). For ν = 0, (9.3.18) and (9.3.19) give respectively (9.3.15) and<br />

(9.3.16). As usual, we are interested in finding the solution U ∈ X of (9.3.4).<br />

W<strong>it</strong>hout entering into details of the proof, we claim that <strong>it</strong> is possible to find a constant<br />

µ ∗ > 0 such that, for any µ ∈]0,µ ∗ ], the bilinear form Bw in (9.3.18) fulfills the<br />

requirements (9.3.5) and (9.3.6). This result is given in funaro(1988) for the Chebyshev<br />

case and can be easily generalized to other Jacobi weights. This implies existence and<br />

uniqueness of a weak solution U. When f ∈ C0 ( Ī), we can relate U to problem (9.1.5)<br />

by arguing as follows. Considering that ˜ φ ′ ∈ P0 is a constant function, we wr<strong>it</strong>e<br />

(9.3.20)<br />

=<br />

<br />

<br />

Ũ<br />

I<br />

′ φ ˜′ dx = (Ũ(1) − Ũ(−1))˜ φ ′ = (U(1) − U(−1)) ˜ φ ′<br />

U<br />

I<br />

′ φ ˜′ ′<br />

dx = U (1) φ(1) ˜ ′<br />

− U (−1) φ(−1) ˜ −<br />

Therefore, since Ũ ′′ ≡ 0 and ˜ φ(±1) = φ(±1), one has<br />

<br />

(9.3.21) Bw(U,φ) = (−U ′′ + µU)(φ − ˜ φ)w dx +<br />

I<br />

<br />

I<br />

<br />

I<br />

U ′′ ˜ φ dx.<br />

(−U ′′ + µU) ˜ φ dx<br />

+ U ′ (1)φ(1) − U ′ (−1)φ(−1) = Fw(φ), ∀φ ∈ X.<br />

One first uses test functions φ belonging to H 1 0,w(I) (therefore ˜ φ ≡ 0). This implies<br />

the differential equation −U ′′ + µU = f in I. Finally, we eliminate the integrals and<br />

obtain the boundary cond<strong>it</strong>ions.


196 Polynomial Approximation of Differential Equations<br />

As pointed out in section 9.2, the use of ultraspherical weight functions w<strong>it</strong>h ν = 0,<br />

can lead to some trouble in the treatment of the integrals. This is why the bilinear form<br />

Bw has been decomposed into two components. In the one containing the weight<br />

function w, integration by parts is allowed since φ − ˜ φ vanishes at the points x = ±1.<br />

The other component does not contain the weight function and takes into account the<br />

boundary constraints.<br />

Variational formulations also exist for fourth-order equations. For example, we<br />

examine problem (9.1.6). We assume σi = 0, 1 ≤ i ≤ 4. Otherwise, we replace U by<br />

U − s, where s ∈ P3 is given in (7.4.20). Then, for f ∈ L 2 w(I), we define<br />

(9.3.22) Bw(ψ,φ) :=<br />

(9.3.23) Fw(φ) :=<br />

A natural choice for the functional space is<br />

<br />

ψ<br />

I<br />

′′ (φw) ′′ dx, ∀ψ,φ ∈ X,<br />

<br />

I<br />

fφw dx, ∀φ ∈ X.<br />

(9.3.24) X ≡ H 2 0,w(I) := {φ| φ ∈ H 2 w(I), φ(±1) = φ ′ (±1) = 0}.<br />

It is left for the reader to check that Bw and Fw satisfy the hypotheses of Lax-Milgram<br />

theorem, when w is the ultraspherical weight function w<strong>it</strong>h −1 < ν < 1. In particular,<br />

coerciveness is a consequence of lemma 8.2.5. This property provides a unique weak<br />

solution U ∈ X of the problem Bw(U,φ) = Fw(φ), ∀φ ∈ X. In the standard way, one<br />

verifies that U is solution to (9.1.6) whenever f ∈ C0 ( Ī). More properties and results<br />

are discussed in bernardi and maday(1990) and funaro and heinrichs(1990).<br />

In a more advanced setting, we can construct variational formulations of the prob-<br />

lems considered above, where two different spaces are used for the solution and the test<br />

functions. In this case, one defines an appropriate bilinear form B : X × Y → R, and<br />

a linear operator F : Y → R, in order that U ∈ X is solution of the weak problem<br />

B(U,φ) = F(φ), ∀φ ∈ Y. A generalization of theorem 9.3.1 is obtained by modifying<br />

cond<strong>it</strong>ion (9.3.6) as follows:


Ordinary Differential Equations 197<br />

(9.3.25)<br />

⎧<br />

⎨ ∀ψ ∈ X, ∃φ ∈ Y : |B(ψ,φ)| ≥ C2 ψXφY,<br />

⎩<br />

∀φ ∈ Y φ ≡ 0, ∃ψ ∈ X : B(ψ,φ) = 0.<br />

Extensions of this kind are studied, for instance, in nečas(1962). This approach was<br />

used in canuto and quarteroni(1984) to provide an alternative variational formula-<br />

tion to problem (9.1.5).<br />

9.4 Approximation of problems in the weak form<br />

Let us see how to use the results of the previous section to carry out the convergence<br />

analysis for the approximation schemes (9.2.14) and (9.2.15). If Bw and Fw are<br />

respectively given by (9.3.2) and (9.3.3) ( X ≡ H 1 0,w(I)), we already noticed that (9.3.4)<br />

has a unique weak solution and generalizes problem (9.1.4) w<strong>it</strong>h σ1 = σ2 = 0. We also<br />

wr<strong>it</strong>e (9.2.14) in a variational form.<br />

For any n ≥ 2, we consider the fin<strong>it</strong>e dimensional space P 0 n ⊂ X (see (6.4.11)). We<br />

remark that P 0 n is a Hilbert space w<strong>it</strong>h the same inner product defined in X. Then,<br />

we introduce the linear operator Fw,n : P 0 n → R as follows:<br />

(9.4.1) Fw,n(φ) :=<br />

where f ∈ L 2 w(I).<br />

<br />

(Πw,n−2f)φw dx, ∀φ ∈ P<br />

I<br />

0 n,<br />

For the sake of simplic<strong>it</strong>y, let w be the ultraspherical weight function w<strong>it</strong>h ν := α = β<br />

satisfying −1 < ν < 1. W<strong>it</strong>h the help of theorem 9.3.2, we can apply theorem 9.3.1 to<br />

obtain a unique solution pn ∈ P 0 n of the problem<br />

(9.4.2) Bw(pn,φ) = Fw,n(φ), ∀φ ∈ P 0 n,


198 Polynomial Approximation of Differential Equations<br />

where Bw has been introduced in (9.3.2). We note that the space in which we seek<br />

the solution is rather small now. On the other hand, we require (9.4.2) to be satisfied<br />

only for test functions belonging to a subspace of X.<br />

It is very easy to check that pn is solution to (9.2.14) where σ1 = σ2 = 0. In fact,<br />

setting a(x) := (1 −x 2 )w(x), x ∈ I, and recalling that p ′′ n +Πw,n−2f ∈ Pn−2, one has<br />

−<br />

<br />

p<br />

I<br />

′′ nφw dx = Fw,n(φ), ∀φ ∈ P 0 n<br />

<br />

− p<br />

I<br />

′′ nχa dx =<br />

<br />

⇕<br />

(Πw,n−2f)χa dx, ∀χ ∈ Pn−2<br />

I<br />

⇕<br />

−Πa,n−2(p ′′ n) = Πa,n−2(Πw,n−2f) in I<br />

⇕<br />

− p ′′ n = Πw,n−2f in I.<br />

There are several reasons to prefer the new formulation (9.4.2). First of all, inequal<strong>it</strong>y<br />

(9.3.9) yields<br />

(9.4.3) p ′ nw ≤ C4 sup<br />

≤ C4 Πw,n−2fw sup<br />

φ∈P 0 n<br />

φ≡0<br />

φw<br />

φ ′ w<br />

φ∈P 0 n<br />

φ≡0<br />

Fw,n(φ)<br />

φ ′ w<br />

≤ C5 f L 2 w (I), ∀n ≥ 2.<br />

To get (9.4.3), we used the Schwarz inequal<strong>it</strong>y, relation (6.2.7) and inequal<strong>it</strong>y (5.7.4).<br />

According to the usual terminology, (9.4.3) is known as a stabil<strong>it</strong>y cond<strong>it</strong>ion. Namely,<br />

a certain norm of the solution pn is bounded independently of n. In add<strong>it</strong>ion, pn<br />

converges to the solution U of (9.3.4) when n tends to infin<strong>it</strong>y. The proof is simple,<br />

being byproduct of a very general result (see strang(1972)).


Ordinary Differential Equations 199<br />

Theorem 9.4.1 - Let X be a Hilbert space. Let B : X × X → R be a bilinear form<br />

and F : X → R be a linear operator, both satisfying the hypotheses of theorem 9.3.1.<br />

Let us denote by U ∈ X the unique solution of the problem B(U,φ) = F(φ), ∀φ ∈ X.<br />

For any n ∈ N, let Xn ⊂ X be a fin<strong>it</strong>e dimensional subspace of dimension n and<br />

Fn : Xn → R be a linear operator. Let us denote by pn ∈ Xn the solution of the<br />

problem B(pn,φ) = Fn(φ), ∀φ ∈ Xn.<br />

Then, we can find a constant C > 0 such that, for any n ∈ N, we have<br />

(9.4.4) U − pnX ≤ C<br />

<br />

inf<br />

χ∈Xn<br />

U − χX + sup<br />

φ∈Xn<br />

φ≡0<br />

Proof - For any n ∈ N, we define χn ∈ Xn such that<br />

(9.4.5) B(χn,φ) = B(U,φ), ∀φ ∈ Xn.<br />

<br />

|F(φ) − Fn(φ)|<br />

.<br />

φX<br />

Existence and uniqueness of χn are obtained by applying theorem 9.3.1, after noting<br />

that B(U, ·) : X → R is a linear operator for any fixed U ∈ X. We use the triangle<br />

inequal<strong>it</strong>y to wr<strong>it</strong>e<br />

(9.4.6) U − pnX ≤ U − χnX + χn − pnX, n ∈ N.<br />

Due to (9.3.6) and (9.4.5), for any χ ∈ Xn, we get<br />

(9.4.7) U − χn 2 X ≤ C −1<br />

2 B(U − χn,U − χn)<br />

= C −1<br />

2 B(U − χn,U − χ) ≤ C1C −1<br />

2 U − χnX U − χX,<br />

which gives an estimate of the error U − χn, n ∈ N.<br />

On the other hand, (9.3.6) yields<br />

(9.4.8) χn − pnX ≤ B(χn − pn,χn − pn)<br />

C2 χn − pnX<br />


200 Polynomial Approximation of Differential Equations<br />

≤ C −1<br />

2 sup<br />

φ∈Xn<br />

φ≡0<br />

B(χn − pn,φ)<br />

φX<br />

= C −1<br />

2 sup<br />

φ∈Xn<br />

φ≡0<br />

|F(φ) − Fn(φ)|<br />

, n ∈ N.<br />

φX<br />

Inequal<strong>it</strong>y (9.4.4) follows by combining (9.4.6), (9.4.7) and (9.4.8).<br />

Returning to problem (9.4.2), we apply the previous theorem to get the estimate<br />

<br />

(9.4.9) U −pnH1 0,w (I) ≤ C<br />

inf<br />

χ∈P0 U −χH1 0,w (I) + f −Πw,n−2fL2 w (I)<br />

n<br />

<br />

, ∀n ≥ 2.<br />

We further bound this error using the results of chapter six. We can consider inequal<strong>it</strong>ies<br />

(6.4.15) and (6.4.6) after taking χ := Π 1 0,w,nU in (9.4.9). We remark that f = −U ′′ in<br />

I. Thus, we finally get<br />

(9.4.10) U − pn H 1 0,w (I) ≤ C<br />

k 1<br />

fH k<br />

w n<br />

(I), ∀n > k ≥ 0.<br />

This shows that pn converges to U w<strong>it</strong>h a rate depending on the regular<strong>it</strong>y of f. Due<br />

to inequal<strong>it</strong>y (5.7.4), this convergence is uniform.<br />

We remark that the function χn, n ≥ 2, defined in the proof of theorem 9.4.1, here<br />

coincides w<strong>it</strong>h the polynomial ˆ Π 1 0,w,nU ∈ P 0 n, introduced in (6.4.13). The existence of<br />

such a polynomial follows from theorem 9.3.1, and estimates of the error U − ˆ Π 1 0,w,nU<br />

can be recovered from (9.4.7).<br />

The same kind of analysis applies to the collocation method (9.2.15) in the case<br />

σ1 = σ2 = 0. It is sufficient to recall the quadrature formula (3.5.1) and modify (9.4.1)<br />

according to<br />

(9.4.11) Fw,n(φ) :=<br />

n<br />

j=0<br />

f(η (n)<br />

j )φ(η (n)<br />

j ) ˜w (n)<br />

j , ∀φ ∈ P 0 n,<br />

where f ∈ C0 ( Ī). We recall the Schwarz inequal<strong>it</strong>y and theorem 3.8.2 to check (9.3.7).<br />

Actually, one has |Fw,n(φ)| ≤ C3f C 0 (Ī)φ H 1 0,w (I), ∀φ ∈ P 0 n, ∀n ≥ 2.


Ordinary Differential Equations 201<br />

Choosing as test functions the Lagrange polynomials φ := ˜l (n)<br />

i , 1 ≤ i ≤ n − 1, we<br />

obtain the relations −p ′′ n(η (n)<br />

i ) = f(η (n)<br />

i ), 1 ≤ i ≤ n − 1 (see (9.2.15) and (7.4.9)).<br />

To show that pn converges to U when n → +∞, we use theorem 9.4.1. Therefore, we<br />

must bound the error |Fw(φ) − Fw,n(φ)|, φ ∈ P 0 n. Recalling relation (3.8.15) one gets<br />

(9.4.12)<br />

<br />

Fw(φ) − Fw,n(φ) = (f − Ĩw,nf,φ)w + [( Ĩw,nf,φ)w − ( Ĩw,nf,φ)w,n] <br />

≤ f − Ĩw,nfwφw<br />

<br />

<br />

un<br />

+ <br />

<br />

2 w,n − un2 w<br />

un2 (<br />

w,n<br />

Ĩw,nf,un)w,n<br />

<br />

<br />

(φ,un)w,n <br />

<br />

unw,n unw,n <br />

<br />

≤ f − Ĩw,nfw + C |(Ĩw,nf,un)w,n|<br />

<br />

φw, ∀φ ∈ P 0 n,<br />

unw,n<br />

where un := P (α,β)<br />

n , n ∈ N, and C > 0 does not grow w<strong>it</strong>h n (see section 3.8). By<br />

formula (3.5.1) and orthogonal<strong>it</strong>y, one has (Πw,n−1ψ,un)w,n = (Πw,n−1ψ,un)w = 0,<br />

∀ψ ∈ Pn. This implies that<br />

(9.4.13)<br />

≤ γ −1<br />

1<br />

|( Ĩw,nf,un)w,n|<br />

unw,n<br />

≤ Ĩw,nf − Πw,n−1( Ĩw,nf)w,n<br />

<br />

Ĩw,nf − fw + f − Πw,n−1fw + Πw,n−1(f − Ĩw,nf)w<br />

<br />

, n ≥ 2.<br />

For the last inequal<strong>it</strong>y we used (3.8.6) and (2.1.5). By combining (9.4.12) and (9.4.13),<br />

final estimates are obtained using the results of sections 6.2 and 6.6. The results of a<br />

numerical experiment are documented in table 11.2.1.<br />

The study of the approximation of elliptic problems, on the basis of their variational<br />

formulations, is a standard technique in fin<strong>it</strong>e element or fin<strong>it</strong>e-differences methods. Cel-<br />

ebrated examples are considered for instance in aubin (1972), strang and fix(1973),<br />

ciarlet(1978), fletcher(1984), and in many other books. The simplest approxima-<br />

tion technique is in general represented by the so called Galerkin method. W<strong>it</strong>hin the<br />

framework of polynomial approximations, this procedure consists in finding pn ∈ P 0 n<br />

such that


202 Polynomial Approximation of Differential Equations<br />

(9.4.14) Bw(pn,φ) = Fw(φ), ∀φ ∈ P 0 n,<br />

where Bw and Fw are respectively given in (9.3.2) and (9.3.3). The convergence<br />

analysis is straightforward, since now the linear operator on the right-hand side does<br />

not depend on n. In add<strong>it</strong>ion, (9.4.14) is equivalent to the problem<br />

(9.4.15)<br />

⎧<br />

⎨<br />

−p ′′ n = Πa,n−2f in I,<br />

⎩<br />

pn(±1) = 0,<br />

where a(x) := (1 − x 2 )w(x), x ∈ I. The solution in this particular case has the<br />

form pn = ˆ Π 1 0,w,nU, ∀n ≥ 2 (see (6.4.13)). However, the linear system associated to<br />

(9.4.15) is not easy to recover when solving the problem in the frequency space relative<br />

to the weight function w. Usually, this system is obtained by testing (9.4.14) on the set<br />

of polynomials: φ(x) := P (α,β)<br />

k<br />

(x) − 1<br />

2<br />

(1 − x)P (α,β)<br />

k<br />

(−1) − 1<br />

2<br />

(α,β)<br />

(1 + x)P k (1), x ∈ Ī,<br />

2 ≤ k ≤ n. By the orthogonal<strong>it</strong>y of Jacobi polynomials, we obtain the corresponding<br />

(n − 1) × (n − 1) matrix which does not explic<strong>it</strong>ly contain the rows relative to the<br />

boundary cond<strong>it</strong>ions as in the tau method (clarifying remarks about this point are<br />

given in boyd(1989), chapter six).<br />

Let us consider now the approximation of other differential equations, such as<br />

(9.1.5). For X ≡ H 1 w(I), a variational formulation is given by (9.3.4), where Bw and<br />

Fw are defined in (9.3.18) and (9.3.19) respectively. Polynomial approximations of the<br />

solution U can be determined in several ways. We leave the analysis of tau and Galerkin<br />

methods to the reader. We examine the collocation method.<br />

We require pn ∈ Pn, n ≥ 2, to be solution of<br />

(9.4.16) Bw,n(pn,φ) = Fw,n(φ), ∀φ ∈ Pn.<br />

Here, Bw,n : Pn × Pn → R and Fw,n : Pn → R are respectively defined by


Ordinary Differential Equations 203<br />

(9.4.17) Bw,n(ψ,φ) :=<br />

<br />

(ψ −<br />

I<br />

˜ ψ) ′ [(φ − ˜ φ)w] ′ dx +<br />

<br />

+ µ (Ia,n−1ψ)(φ −<br />

I<br />

˜ <br />

φ)w dx + µ<br />

(9.4.18) Fw,n(φ) :=<br />

<br />

(Ia,n−1f)(φ −<br />

I<br />

˜ φ)w dx +<br />

<br />

I<br />

˜ψ ′ ˜ φ ′ dx<br />

(Ia,n−1ψ)<br />

I<br />

˜ φ dx, ∀ψ,φ ∈ Pn,<br />

+σ2φ(1) − σ1φ(−1), ∀φ ∈ Pn,<br />

<br />

(Ia,n−1f)<br />

I<br />

˜ φ dx<br />

where a(x) := (1 − x 2 )w(x), x ∈ I, and the interpolation operator Ia,n−1, n ≥ 2, is<br />

defined in section 3.3.<br />

Our aim is to show that (9.4.16) leads to a collocation type approximation. First,<br />

assume φ ∈ P 0 n so that ˜ φ ≡ 0. W<strong>it</strong>h the help of integration by parts and quadrature<br />

formula (3.5.1), we obtain<br />

(9.4.19) −<br />

n<br />

j=0<br />

=<br />

p ′′ n(η (n)<br />

j )φ(η (n)<br />

j ) ˜w (n)<br />

j<br />

n<br />

j=0<br />

+ µ<br />

n<br />

j=0<br />

(Ia,n−1pn)(η (n)<br />

j )φ(η (n)<br />

j ) ˜w (n)<br />

j<br />

(Ia,n−1f)(η (n)<br />

j )φ(η (n)<br />

j ) ˜w (n)<br />

j , ∀φ ∈ P 0 n.<br />

Choosing φ := ˜ l (n)<br />

i , 1 ≤ i ≤ n − 1, (9.4.19) implies<br />

(9.4.20) −p ′′ n(x) + µ(Ia,n−1pn)(x) = (Ia,n−1f)(x) at x = η (n)<br />

i , 1 ≤ i ≤ n − 1.<br />

Note that, for any g ∈ C0 ( Ī), one has<br />

(n)<br />

(n)<br />

(Ia,n−1g)(η i ) = ( Ĩw,ng)(η i ) = g(η (n)<br />

i ),<br />

1 ≤ i ≤ n − 1. This gives −p ′′ n(η (n)<br />

i ) + µpn(η (n)<br />

i ) = f(η (n)<br />

i ), 1 ≤ i ≤ n − 1. Moreover,<br />

for a general φ ∈ Pn, we have a relation like (9.3.21). Both −p ′′ n + µIa,n−1pn and<br />

Ia,n−1f are polynomials of degree n −2. Since they coincide at n −1 points, (9.4.20)<br />

holds for any x ∈ Ī. This enables us to remove the integrals and recover the boundary<br />

cond<strong>it</strong>ions −p ′ n(−1) = −σ1 and p ′ n(1) = σ2. To obtain the corresponding linear<br />

system we follow the suggestions of section 7.4 (see (7.4.13)).


204 Polynomial Approximation of Differential Equations<br />

To develop the convergence analysis, an extension of theorem 9.4.1 is needed, since<br />

the bilinear form now depends on n. For the sake of simplic<strong>it</strong>y, we do not discuss<br />

this generalization. The reader finds indications to proceed in this investigation in<br />

strang(1972) and ciarlet(1978).<br />

A different treatment of the boundary cond<strong>it</strong>ions can be taken into consideration.<br />

First, we examine the Legendre case (w ≡ 1). In (9.4.16) define Bw,n : Pn × Pn → R<br />

and Fw,n : Pn → R such that<br />

(9.4.21) Bw,n(ψ,φ) :=<br />

(9.4.22) Fw,n(φ) :=<br />

n<br />

j=0<br />

<br />

n<br />

ψ<br />

I<br />

′ φ ′ dx + µ<br />

j=0<br />

ψ(η (n)<br />

j )φ(η (n)<br />

j ) ˜w (n)<br />

j , ∀ψ,φ ∈ Pn,<br />

f(η (n)<br />

j )φ(η (n)<br />

j ) ˜w (n)<br />

j + σ2φ(1) − σ1φ(−1), ∀φ ∈ Pn.<br />

We replaced the integrals in (9.3.15) and (9.3.16) by the summations. The modification<br />

does not alter the scheme at the nodes inside the interval I. In fact, using the test<br />

functions φ := ˜l (n)<br />

i , 1 ≤ i ≤ n − 1, we still get the set of equations −p ′′ n(η (n)<br />

i ) +<br />

µpn(η (n)<br />

i ) = f(η (n)<br />

i ), 1 ≤ i ≤ n − 1. At the boundary points one gets<br />

(9.4.23)<br />

⎧<br />

⎨ −p ′′ n(−1) + µpn(−1) − γp ′ n(−1) = f(−1) − γσ1,<br />

⎩<br />

−p ′′ n(1) + µpn(1) + γp ′ n(1) = f(1) + γσ2,<br />

where γ := [ ˜w (n)<br />

0 ] −1 = [ ˜w (n)<br />

n ] −1 = 1<br />

2<br />

n(n + 1). Thus, we obtained the system (7.4.16)<br />

w<strong>it</strong>h q := Ĩw,nf. Note that in (9.4.20) the equation is also collocated at the points<br />

x = ±1, and the boundary cond<strong>it</strong>ions are used as corrective terms.<br />

We can make an interesting remark. In the previous examples, the variational<br />

formulations have been presented as an intermediate step to obtain convergence results<br />

by virtue of general abstract theorems. Here, we followed an inverse path. Starting<br />

from an appropriate weak formulation, we recovered a new approximation scheme w<strong>it</strong>h<br />

different cond<strong>it</strong>ions at the boundary. These relations may not have a direct physical<br />

interpretation. Nevertheless, convergence is still achieved. As a matter of fact, one<br />

shows that limn→+∞ p ′ n(−1) = σ1, limn→+∞ p ′ n(1) = σ2 (see funaro(1988)).


Ordinary Differential Equations 205<br />

There are several reasons for preferring cond<strong>it</strong>ions (9.4.23) to exact boundary condi-<br />

tions. For instance, as noticed in section 8.6, the eigenvalues of the matrix corresponding<br />

to system (7.4.16) are more appropriate for computation than the eigenvalues of problem<br />

(8.6.8). In add<strong>it</strong>ion, for the approximation of the solution of (9.1.5), the new approach<br />

gives better numerical results. We compare the two ways of imposing the boundary<br />

cond<strong>it</strong>ions in table 9.4.1. Take µ := 1 and f, σ1, σ2 such that the exact solution of<br />

(9.1.5) is U(x) := cos(e x ), x ∈ Ī. Then, compute the error En := pn − Ĩw,nUw<br />

for various n. In the first column, pn is obtained by the collocation method w<strong>it</strong>h the<br />

cond<strong>it</strong>ions p ′ n(−1) = σ1, p ′ n(1) = σ2. In the second column, pn is obtained by the<br />

collocation method w<strong>it</strong>h the cond<strong>it</strong>ions (9.4.23).<br />

n Column 1 Column 2<br />

6 .1090 × 10 −1 .2009 × 10 −3<br />

8 .4130 × 10 −2 .1840 × 10 −4<br />

10 .2723 × 10 −3 .5941 × 10 −6<br />

12 .6896 × 10 −5 .1495 × 10 −7<br />

14 .1404 × 10 −5 .1471 × 10 −8<br />

16 .7122 × 10 −7 .4405 × 10 −10<br />

Table 9.4.1 - Errors corresponding to the Legendre<br />

collocation approximation of problem (9.1.5).<br />

For other Jacobi ultraspherical weights, we generalize (9.4.21) and (9.4.22) by defining<br />

(9.4.24) Bw,n(ψ,φ) :=<br />

+ µ<br />

n<br />

j=0<br />

ψ(η (n)<br />

j ) (φ − ˜ φ)(η (n)<br />

j ) ˜w (n)<br />

j<br />

<br />

(ψ −<br />

I<br />

˜ ψ) ′ [(φ − ˜ φ)w] ′ dx +<br />

+ µ<br />

n<br />

j=0<br />

<br />

I<br />

˜ψ ′ ˜ φ ′ dx<br />

ψ(η (n)<br />

j ) ˜ φ(η (n)<br />

j ) χ (n)<br />

j , ∀ψ,φ ∈ Pn,


206 Polynomial Approximation of Differential Equations<br />

(9.4.25) Fw,n(φ) :=<br />

+ µ<br />

n<br />

j=0<br />

n<br />

j=0<br />

f(η (n)<br />

j ) (φ − ˜ φ)(η (n)<br />

j ) ˜w (n)<br />

j<br />

f(η (n)<br />

j ) ˜ φ(η (n)<br />

j ) χ (n)<br />

j + σ2φ(1) − σ1φ(−1), ∀φ ∈ Pn.<br />

The weights χ (n)<br />

j , 0 ≤ j ≤ n, are defined in section 3.7. When ν = 0 , we obtain<br />

(9.4.21) and (9.4.22) by noting that χ (n)<br />

j<br />

= ˜w(n)<br />

j , 0 ≤ j ≤ n.<br />

Basically, Bw,n and Fw,n are derived from (9.3.18) and (9.3.19) respectively, by<br />

considering quadrature formulas in place of integrals. The errors introduced decay<br />

spectrally, giving the convergence of the method. We remark that, although Clenshaw-<br />

Curtis formula does not have the accuracy of Gaussian formulas, <strong>it</strong> is only used for<br />

polynomials of low degree (we have ˜ φ ∈ P1 and ψ ˜ φ ∈ Pn+1).<br />

The corresponding variational problem is equivalent to system (7.4.16) where we set<br />

q := Ĩw,nf and γ := [χ (n)<br />

0 ]−1 = [χ (n)<br />

n ] −1 . In general, if γ grows at least as n 2 ,<br />

one gets convergence of pn to the solution of (9.1.5) when n tends to infin<strong>it</strong>y (see<br />

funaro(1988)).<br />

The idea of collocating the equation at the boundary points has been also proposed<br />

in funaro and gottlieb(1988) for first-order equations, such as (9.1.1). This leads<br />

to a collocation scheme like that in (7.4.4), where γ is proportional to n 2 . A proof<br />

of convergence is easily given for the Legendre case following the guideline of theorem<br />

9.2.1.<br />

The techniques introduced and analyzed in this section are also used for the approx-<br />

imation of the solution of fourth-order equations. The collocation method for computing<br />

the solution of problem (9.1.6) is<br />

⎧<br />

(9.4.26)<br />

p IV<br />

n (η (n)<br />

i ) = f(η (n)<br />

i ), 1 ≤ i ≤ n − 1,<br />

⎪⎨<br />

pn(−1) = σ1, pn(1) = σ2,<br />

⎪⎩<br />

p ′ n(−1) = σ3, p ′ n(1) = σ4,<br />

where pn is a polynomial in Pn+2, n ≥ 2.


Ordinary Differential Equations 207<br />

The way to wr<strong>it</strong>e the linear system associated w<strong>it</strong>h (9.4.26) is described in section 7.4.<br />

A proof of convergence for ultraspherical weights (ν := α = β, −1 < ν < 1) is provided<br />

in bernardi and maday(1990). In the same paper, another collocation scheme is<br />

proposed and analyzed. In the latter approach the nodes η (n)<br />

i , 1 ≤ i ≤ n − 1, in<br />

(9.4.26) are replaced by the nodes ξ (n−1)<br />

i , 1 ≤ i ≤ n − 1, which are the zeroes of<br />

P (ν+2,ν+2)<br />

n−1 , n ≥ 2, −1 < ν < 1. In both cases, error estimates in the space H2 0,w(I)<br />

(see (9.3.24)) are given. The theory is developed w<strong>it</strong>h the help of quadrature formulas<br />

having a double multiplic<strong>it</strong>y at the boundary points (see also gautschi and li(1991)).<br />

Other boundary cond<strong>it</strong>ions are analyzed in funaro and heinrichs(1990).<br />

9.5 Approximation in unbounded domains<br />

In this section, we discuss examples where the domain I ⊆ R is not bounded. Let us<br />

assume I :=]0,+∞[. A typical second-order problem consists in finding U : I → R<br />

such that<br />

(9.5.1)<br />

⎧<br />

⎨ −U ′′ + µU = g in I,<br />

⎩<br />

U(0) = σ,<br />

where σ ∈ R, µ > 0, and g : I → R are given.<br />

In order to determine a unique solution of (9.5.1), we need an extra cond<strong>it</strong>ion. This<br />

is obtained by prescribing the behavior of the unknown U at infin<strong>it</strong>y. For instance, we<br />

can require the convergence of U to a given value. W<strong>it</strong>hout loss of general<strong>it</strong>y we can<br />

impose the cond<strong>it</strong>ion limx→+∞ U(x) = 0. To develop the analysis, we make further<br />

assumptions on the rate of decay of U. This is obtained by assuming that U belongs to<br />

some functional space X. Thus, the best way to argue is to use a su<strong>it</strong>able variational<br />

formulation of (9.5.1). We remark that <strong>it</strong> is not restrictive to assume that σ = 0.<br />

Let v(x) := x α e x , x ∈ I, −1 < α < 1, be a weight function. Let us define the Hilbert<br />

space:


208 Polynomial Approximation of Differential Equations<br />

(9.5.2) X ≡ H 1 0,v(I) := {φ |φ ∈ H 1 v(I), φ(0) = 0}.<br />

According to section 5.7, the norm in X is given by φX := <br />

I φ2vdx + <br />

∀φ ∈ X. Then, for g ∈ L2 v(I), we are concerned w<strong>it</strong>h finding U ∈ X such that<br />

(9.5.3)<br />

+∞<br />

U<br />

0<br />

′ (φv) ′ +∞<br />

dx + µ<br />

0<br />

Uφv dx =<br />

+∞<br />

0<br />

I [φ′ ] 2 vdx<br />

gφv dx, ∀φ ∈ X.<br />

W<strong>it</strong>h the same arguments of section 9.3, using the Lax-Milgram theorem, we get exis-<br />

tence and uniqueness of a weak solution of (9.5.3), provided the parameter µ is larger<br />

than 1<br />

2 max{1, 1<br />

1−α }. The crucial part of the proof is to show coerciveness (see (9.3.6)).<br />

This follows by virtue of lemma 8.2.7.<br />

One checks that U ∈ X implies that limx→+∞ U(x) = 0, and the decay is exponential.<br />

Therefore, due to theorem 6.1.4, we have the requis<strong>it</strong>es to approximate the solution of<br />

problem (9.5.3) by Laguerre functions (see section 6.7). To this end, we introduce the<br />

subspace S 0 n := {φ| φ ∈ Sn, φ(0) = 0} ⊂ X.<br />

For example, when g ∈ C0 ( Ī), we discretize (9.5.3) as follows. For any n ≥ 2, we seek<br />

Pn ∈ S0 n−1 such that<br />

+∞<br />

(9.5.4)<br />

0<br />

P ′ n(φv) ′ +∞<br />

dx + µ Pnφv dx =<br />

0<br />

+∞<br />

0<br />

( Ĩ∗ v,ng)φv dx, ∀φ ∈ S 0 n−1.<br />

The interpolation operator Ĩ∗ v,n : C 0 ( Ī) → Sn−1, n ≥ 2, is defined by the relation<br />

Ĩ ∗ v,ng := [ Ĩw,n(ge x )]e −x , ∀g ∈ C 0 ( Ī) (see also section 6.7), where Ĩw,n, n ≥ 1, is the<br />

Laguerre Gauss-Radau interpolation operator (see section 3.3). We now apply theorem<br />

9.4.1. Estimating the error g − Ĩ∗ v,ng L 2 v (I), one deduces that Pn converges to U, as n<br />

tends to infin<strong>it</strong>y, in the norm of the space X. The convergence is also uniform because<br />

of the inequal<strong>it</strong>y<br />

(9.5.5) sup<br />

x∈I<br />

|φ(x)e x/2 | ≤ C φX, ∀φ ∈ X.<br />

The next step is to determine the solution to (9.5.4). We consider the subst<strong>it</strong>utions:<br />

pn(x) := Pn(x)ex , x ∈ Ī, n ≥ 2, and f(x) := g(x)ex , x ∈ Ī. Therefore, integration by<br />

parts allows us to wr<strong>it</strong>e (note that the boundary terms are vanishing):<br />

1<br />

2 ,


Ordinary Differential Equations 209<br />

(9.5.6)<br />

+∞<br />

(−p<br />

0<br />

′′ n + 2p ′ +∞<br />

n − pn)φw dx = (<br />

0<br />

Ĩw,nf)φw dx, ∀φ ∈ P 0 n−1.<br />

In (9.5.6), w(x) := xαe−x , x ∈ I, −1 < α < 1, is the Laguerre weight function. At this<br />

point, we use the quadrature formula (3.6.1). Taking as test functions the Lagrange<br />

polynomials ˜l (n)<br />

i , 1 ≤ i ≤ n − 1, we obtain the collocation scheme<br />

⎧<br />

⎨<br />

(9.5.7)<br />

⎩<br />

−p ′′ n(η (n)<br />

i ) + 2p ′ n(η (n)<br />

i ) + (µ − 1)pn(η (n)<br />

i ) = f(η (n)<br />

i ), 1 ≤ i ≤ n − 1,<br />

pn(η (n)<br />

0 ) = 0,<br />

which is the same one presented in (7.4.26) w<strong>it</strong>h σ = 0, q := Ia,n−1f, and a(x) :=<br />

xw(x), x ∈ I. For the numerical implementation we suggest taking into account the<br />

results of sections 3.10 and 7.5.<br />

Other differential equations are approximated in a similar way. Examples are found<br />

in maday, pernaud-thomas and vandeven(1985), and mavriplis(1989). For the<br />

domain I ≡ R, Herm<strong>it</strong>e functions are a su<strong>it</strong>able basis for computations when the<br />

solution decays like e −x2<br />

kavian(1988).<br />

at infin<strong>it</strong>y. Results in this direction are given in funaro and<br />

The l<strong>it</strong>erature offers several other numerical techniques for the treatment of prob-<br />

lems in unbounded domains by spectral methods. For instance, one can map the do-<br />

main I in the interval ] −1,1[ and approximate the new problem by Jacobi polynomial<br />

expansions. The main difficulty is to find coordinate transformations preserving the<br />

smoothness of the problem to avoid singular<strong>it</strong>ies and retain spectral accuracy. Early<br />

results were given in grosch and orszag(1977). Further indications and comparisons<br />

are provided in boyd(1982) and boyd(1989), chapter 13. Using the mapping approach,<br />

the solution of the original problem is not required to decay exponentially at infin<strong>it</strong>y as<br />

in the case of Laguerre or Herm<strong>it</strong>e approximations.<br />

The idea of truncating the domain is often used in applications. Here, <strong>it</strong> is not<br />

easy in general to detect the appropriate size of the computational domain. In add<strong>it</strong>ion,<br />

one introduces artificial boundaries, where a non correct specification of the behavior of<br />

the approximating polynomials may cause loss in accuracy. We describe an example in<br />

section 12.2.


210 Polynomial Approximation of Differential Equations<br />

9.6 Other techniques<br />

Variations of the principal methods exposed in the previous sections have been sug-<br />

gested by different authors, w<strong>it</strong>h the aim of preserving spectral accuracy and facil<strong>it</strong>ate<br />

numerical implementation.<br />

For example, we already noted that discretizations by the collocation method of lin-<br />

ear differential operators w<strong>it</strong>h non-constant coefficients, are much easier to implement<br />

than those using tau method. The discrepancy is more remarkable for nonlinear equa-<br />

tions (see sections 3.3 and 9.8). However, we can still work in the frequency space by<br />

combining different techniques. For instance, we modify problem (9.2.6) by setting<br />

(9.6.1)<br />

⎧<br />

⎨p<br />

⎩<br />

′ n + Πw,n−1[ Ĩw,n(Apn)] = Πw,n−1f<br />

pn(−1) = σ.<br />

in ] − 1,1],<br />

We can solve the linear system corresponding to (9.6.1) by an <strong>it</strong>erative method. In this<br />

way, whenever we need to evaluate the product Apn , we can go to the physical space<br />

by Fourier transform (see section 4.1) and perform this operation therein. We return to<br />

the frequency space by the inverse transform. Calculations are faster in the Chebyshev<br />

case, where we can use the FFT (see section 4.3). We remark that this algor<strong>it</strong>hm does<br />

not coincide w<strong>it</strong>h a genuine tau method. The computations, however, exhib<strong>it</strong> a spectral<br />

accuracy.<br />

Other methods have been developed starting from variational techniques. We con-<br />

sider for instance problem (9.1.4), which has a weak formulation given by (9.3.2)-(9.3.3)-<br />

(9.3.4) for σ1 = σ2 = 0. The Galerkin approximation is presented in (9.4.14) which<br />

is equivalent to (9.4.15). Assuming that w ≡ 1, from the orthogonal<strong>it</strong>y of Legendre<br />

polynomials, we easily wr<strong>it</strong>e the corresponding linear system, where the unknowns are<br />

the Fourier coefficients of pn, n ≥ 2. Using the same weight function, a different choice<br />

of the polynomial basis leads to the so called spectral element method introduced in<br />

patera(1984) (see also section 11.2). In this method, pn is expanded in terms of<br />

Chebyshev polynomials, i.e. pn = n<br />

k=0 ckTk. Therefore, subst<strong>it</strong>uting in (9.4.14), one<br />

obtains


Ordinary Differential Equations 211<br />

(9.6.2)<br />

n<br />

k=0<br />

1<br />

ck T<br />

−1<br />

′ k φ ′ dx =<br />

1<br />

−1<br />

fφ dx, ∀φ ∈ P 0 n.<br />

The coefficients ck, 0 ≤ k ≤ n, are determined from the boundary cond<strong>it</strong>ions (these<br />

imply n k=0 ck = n k=0 (−1)kck = 0) and testing (9.6.2) on the set of polynomials<br />

φ(x) := Tj(x) − 1<br />

(1 + x), x ∈ Ī, 2 ≤ j ≤ n. The matrix corresponding<br />

2 (−1)j (1 − x) − 1<br />

2<br />

to the linear system is full and <strong>it</strong>s entries are obtained by computing the integrals<br />

Ikj := 1<br />

−1 T ′ kT ′ jdx. We can provide an explic<strong>it</strong> expression of these quant<strong>it</strong>ies by recalling<br />

(1.5.7) and arguing as in section 2.6. In this approach, we can combine the benef<strong>it</strong>s of<br />

working w<strong>it</strong>h Chebyshev polynomials, w<strong>it</strong>h a variational formulation which does not<br />

involve singular weight functions. The collocation scheme is instead obtained by taking<br />

pn = n−1 (n)<br />

j=1 pn(η j ) ˜l (n)<br />

j<br />

9.7 Boundary layers<br />

in (9.4.14), and testing on φ := ˜ l (n)<br />

j , 1 ≤ j ≤ n − 1.<br />

Small perturbations of certain differential operators can force the solution to degen-<br />

erate in a singular behavior near the boundary points, which is often recognized as a<br />

boundary layer. The terminology was first introduced in prandtl(1905). General mo-<br />

tivations and theoretical results are given in eckhaus(1979), chang and howes(1984),<br />

lagerstrom(1988). A typical example is given by the equation<br />

(9.7.1)<br />

⎧<br />

⎨<br />

−ǫ U ′′<br />

ǫ + Uǫ = 0 in I :=] − 1,1[,<br />

⎩<br />

Uǫ(−1) = 0, Uǫ(1) = 1,<br />

where ǫ > 0 is a given parameter. The solution of (9.7.1) is<br />

(9.7.2) Uǫ(x) = e(1+x)/√ ǫ − e −(x+1)/ √ ǫ<br />

e 2/√ ǫ − e −2/ √ ǫ<br />

, x ∈ Ī.


212 Polynomial Approximation of Differential Equations<br />

When ǫ tends to zero, Uǫ converges to the discontinuous function: U0 ≡ 0 in [−1,1[,<br />

U0(1) = 1. For very small values of ǫ, various physical phenomena are related to the<br />

function Uǫ, which displays sharp derivatives in a neighborhood of the point x = 1. Due<br />

to this behavior, approximations of Uǫ by low degree global polynomials, are affected<br />

by oscillations. Nevertheless, we are not in the s<strong>it</strong>uation that characterizes the Gibbs<br />

phenomenon (see sections 6.2 and 6.8, as well gottlieb and orszag(1977), p.40). Ac-<br />

tually, since the solution is analytic for ǫ > 0, uniform convergence w<strong>it</strong>h an exponential<br />

rate is obtained w<strong>it</strong>h the usual techniques. Tau, Galerkin and collocation methods have<br />

been studied in canuto(1988). In particular, the tau method is equivalent to finding<br />

pǫ,n ∈ Pn, n ≥ 2, ǫ > 0, satisfying problem (7.3.5), where: A ≡ 0, B ≡ 1<br />

ǫ , q ≡ 0,<br />

σ1 = 0, σ2 = 1. Thus, w<strong>it</strong>h the notations of section 7.1, the corresponding linear system<br />

is<br />

(9.7.3)<br />

⎧<br />

⎨<br />

⎩<br />

−ǫ c (2)<br />

k + ck = 0 0 ≤ k ≤ n − 2,<br />

n<br />

k=0 ckuk(−1) = 0, n<br />

k=0 ckuk(1) = 1.<br />

The coefficients c (2)<br />

k , 0 ≤ k ≤ n, are given for ν := α = β in karageorgis and<br />

phillips(1989) (see also (7.1.9) and (7.1.10)). These are all pos<strong>it</strong>ive for ν > −1.<br />

Therefore, <strong>it</strong> is easy to deduce that ck > 0, 0 ≤ k ≤ n. This property is used in<br />

canuto(1988) to estimate the maximum norm of pǫ,n. For instance, from (1.3.9), we<br />

have<br />

(9.7.4) |pǫ,n(x)| =<br />

≤<br />

n<br />

k=0<br />

ck<br />

<br />

n + ν<br />

n<br />

=<br />

<br />

<br />

<br />

<br />

<br />

n<br />

<br />

<br />

<br />

ckuk(x) <br />

≤<br />

k=0<br />

n<br />

ck|uk(x)|<br />

k=0<br />

n<br />

ckuk(1) = pǫ,n(1) = 1, ∀x ∈ Ī, ∀ǫ > 0, ∀n ≥ 2.<br />

k=0<br />

In add<strong>it</strong>ion, using relation (1.4.8), in the Legendre case one gets<br />

(9.7.5) |pǫ,n(x)| ≤<br />

n<br />

k=0<br />

ck|Pk(x)| ≤ 1<br />

2 (1 + x2 )<br />

n<br />

k=0<br />

ckPk(1) = 1<br />

2 (1 + x2 ),<br />

∀x ∈ Ī, ∀ǫ > 0, ∀n ≥ 2.


Ordinary Differential Equations 213<br />

By refining the above inequal<strong>it</strong>ies, we can further investigate the behavior of pǫ,n near<br />

the boundary layer.<br />

In figures 9.7.1 and 9.7.2, we plot the polynomial pǫ,n for n = 16 and n = 22, cor-<br />

responding to the approximation by the tau method of problem (9.7.1) in the Legendre<br />

case (ν = 0) w<strong>it</strong>h ǫ = 10 −4 .<br />

Figure 9.7.1 - Tau-Legendre Figure 9.7.2 - Tau-Legendre<br />

approximation of problem (9.7.1) approximation of problem (9.7.1)<br />

for n = 16 and ǫ = 10 −4 . for n = 22 and ǫ = 10 −4 .<br />

For higher values of n, we have enough resolution to control the violent deviation of Uǫ<br />

near the point x = 1. Then, the oscillations disappear. This happens approximately<br />

when n is larger than C/ √ ǫ, where the constant C > 0 does not depend on ǫ.<br />

Similar conclusions follow for the collocation method. Again, we consider the ul-<br />

traspherical case, where the polynomial pǫ,n, n ≥ 2, ǫ > 0, satisfies:


214 Polynomial Approximation of Differential Equations<br />

(9.7.6)<br />

⎧<br />

⎪⎨<br />

−ǫ p ′′<br />

ǫ,n + pǫ,n =<br />

⎪⎩<br />

pǫ,n(−1) = 0, pǫ,n(1) = 1,<br />

<br />

(n + 2ν) cn−1 cn<br />

+<br />

(n + ν)(2n + 2ν − 1) n x<br />

<br />

u ′ n<br />

in I,<br />

where ck, 0 ≤ k ≤ n, are the Fourier coefficients of pǫ,n and un := P (ν,ν)<br />

n , ν > −1.<br />

In fact, the polynomial u ′ n vanishes at the points η (n)<br />

i , 1 ≤ i ≤ n − 1, giving the<br />

collocation scheme. The corrective term on the right-hand side of the first equation in<br />

(9.7.6) is obtained by equating the coefficients of the monomials x n and x n−1 to the<br />

respective coefficients of the left-hand side. The Chebyshev case is treated in a similar<br />

way.<br />

We can solve (9.7.6) in the frequency space by recalling (2.3.8) and (7.1.3). Due to<br />

the pos<strong>it</strong>iv<strong>it</strong>y of the quant<strong>it</strong>ies involved in these computations, we still recover ck > 0,<br />

0 ≤ k ≤ n. Therefore, inequal<strong>it</strong>ies (9.7.4) and (9.7.5) also hold when pǫ,n is the<br />

approximation associated w<strong>it</strong>h the collocation method.<br />

The same analysis can be developed for the problem<br />

⎧<br />

⎨<br />

(9.7.7)<br />

⎩<br />

Uǫ(−1) = 0, Uǫ(1) = 1.<br />

−ǫ U ′′<br />

ǫ + U ′ ǫ = 0 in I, ǫ > 0,<br />

For this example, the results given in canuto(1988) in the Chebyshev case point out a<br />

different behavior of the approximating polynomials, depending on the par<strong>it</strong>y of their<br />

degree.<br />

9.8 Nonlinear equations<br />

In a large number of practical applications, physics events are modelled by the solution<br />

of nonlinear differential equations. Of course, spectral methods can be employed in


Ordinary Differential Equations 215<br />

these s<strong>it</strong>uations as well, although a lot of robust algor<strong>it</strong>hms are already available in the<br />

framework of fin<strong>it</strong>e-difference computations. The success of a certain approximation<br />

technique is strictly dependent on the type of nonlinear<strong>it</strong>y, so that <strong>it</strong> is difficult to<br />

provide general recipes. Some examples are discussed in chapter twelve. As far as the<br />

theoretical analysis is concerned, references are very few, both for the difficulty and the<br />

modern<strong>it</strong>y of the subject. We give some hints for the implementation. Consider the<br />

problem<br />

(9.8.1)<br />

⎧<br />

⎨U<br />

′ (x) = F(x,U(x)) ∀x ∈] − 1,1],<br />

⎩<br />

U(−1) = σ,<br />

where F : [−1,1] × R → R is a given function and σ ∈ R. We recover the linear<br />

equation (9.1.3) by setting F(x,y) := f(x) − A(x)y.<br />

Appropriate cond<strong>it</strong>ions on F insure existence, uniqueness and regular<strong>it</strong>y of the solu-<br />

tion U (see for instance golomb and shanks (1965), brauer and nohel(1967),<br />

simmons(1972), etc.). According to the remarks of section 3.3, approximations in the<br />

physical space are preferable to approximations in the frequency space, although one<br />

can note a l<strong>it</strong>tle deterioration in the numerical results, due to the effects of the aliasing<br />

errors (see section 4.2 and orszag(1972)). It is common opinion to attribute this phe-<br />

nomenon to the spacing of the collocation grid, when <strong>it</strong> is not adequate to resolve high<br />

frequency oscillations. De-aliasing procedures have been proposed by various authors.<br />

The reader is addressed to boyd(1989) for a collection of references.<br />

The collocation method applied to problem (9.8.1) consists in finding pn ∈ Pn, n ≥ 1,<br />

such that<br />

(9.8.2)<br />

⎧<br />

⎨<br />

⎩<br />

p ′ n(η (n)<br />

i ) = F(η (n)<br />

i ,pn(η (n)<br />

i )) 1 ≤ i ≤ n,<br />

pn(η (n)<br />

0 ) = σ.<br />

In the Legendre case, a convergence result is obtained by reviewing the proof of theorem<br />

9.2.1. This time, we assume that there exists a constant ǫ > 0 such that


216 Polynomial Approximation of Differential Equations<br />

(9.8.3) −[F(x,y1) − F(x,y2)](y1 − y2) ≥ ǫ (y1 − y2) 2 , ∀x ∈ [−1,1], ∀y1,y2 ∈ R.<br />

For a regular F this is equivalent to requiring − ∂F<br />

∂y<br />

≥ ǫ in [−1,1] × R.<br />

It is clear that (9.8.2) is not equivalent to a linear system and the unknown vector<br />

¯v := {pn(η (n)<br />

i )}0≤i≤n has to be computed by an <strong>it</strong>erative approach. We can adopt the<br />

Richardson method (see section 7.6) by defining the sequence ¯v (k) ≡ (v (k)<br />

0 , · · · ,v(k) n ),<br />

k ∈ N, such that<br />

(9.8.4) ¯v (k+1)<br />

:= (I − θD)¯v (k) + θ ¯w (k) , k ∈ N,<br />

where ¯v (0) is the in<strong>it</strong>ial guess. In (9.8.4), D is the matrix corresponding to the system<br />

(7.4.1) and ¯w (k) ≡ σ,F(η (n)<br />

1 ,v (k)<br />

1<br />

), · · · ,F(η(n) n ,v (k)<br />

n ) , k ∈ N. If the <strong>it</strong>erates converge<br />

for some θ > 0, then one gets limk→+∞ v (k)<br />

0 = σ, limk→+∞ v (k) (n)<br />

i = pn(η i ), 1 ≤ i ≤ n.<br />

A faster convergence is realized by using precond<strong>it</strong>ioners (see sections 8.3 and 8.5).<br />

Second-order nonlinear problems can be also considered. An example is given by<br />

the equation<br />

(9.8.5)<br />

⎧<br />

⎨<br />

−U ′′ (x) + F(x,U(x),U ′ (x)) = 0 x ∈] − 1,1[,<br />

⎩<br />

U(−1) = σ1, U(1) = σ2,<br />

where F : [−1,1] × R 2 → R, σ1 ∈ R, σ2 ∈ R, are given.<br />

Spectral type approximations of (9.8.5) are studied in maday and quarteroni(1982)<br />

in the case F(x,y,z) := 1<br />

ǫ yz − f(x), σ1 = σ2 = 0, where ǫ > 0 and the function<br />

f : [−1,1] → R is given. A general trick to analyze the solution of problem (9.8.5) and<br />

<strong>it</strong>s approximation is to wr<strong>it</strong>e the equation in the form U = LF(x,U,U ′ ), where L is<br />

the linear operator which associates to any function f the solution of problem (9.1.4)<br />

w<strong>it</strong>h σ1 = σ2 = 0. Then, the investigation proceeds by using fixed-point theorems and<br />

the compactness of the operator L. More details can be found in brezzi, rappaz and<br />

raviart(1980). We further examine the above example in section 10.4.


Ordinary Differential Equations 217<br />

9.9 Systems of differential equations<br />

Approximation by spectral methods of systems of differential equations in two or more<br />

unknowns can be also taken into consideration. We illustrate a simple example which<br />

will be also useful in section 12.3. Other systems are examined in section 10.5.<br />

We are concerned w<strong>it</strong>h finding the solutions V : [−1,1] → R and W : [−1,1] → R of<br />

the differential problem<br />

(9.9.1)<br />

⎧<br />

−V<br />

⎪⎨<br />

⎪⎩<br />

′′ + µW = f in ] − 1,1[,<br />

−W ′′ − µV = g in ] − 1,1[,<br />

V (±1) = W(±1) = 0,<br />

where f and g are given continuous functions and µ ∈ R.<br />

We approximate (9.9.1) by the collocation method. Thus, for n ≥ 1, we have to find<br />

two polynomials pn,qn ∈ Pn such that<br />

(9.9.2)<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

−p ′′ n(η (n)<br />

i ) + µqn(η (n)<br />

i ) = f(η (n)<br />

i ) 1 ≤ i ≤ n − 1,<br />

−q ′′<br />

n(η (n)<br />

i ) − µpn(η (n)<br />

i ) = g(η (n)<br />

i ) 1 ≤ i ≤ n − 1,<br />

pn(η (n)<br />

0 ) = qn(η (n)<br />

0 ) = 0,<br />

pn(η (n)<br />

n ) = qn(η (n)<br />

n ) = 0.<br />

When n tends to infin<strong>it</strong>y, pn converges to V and qn converges to W. For the proof we<br />

can use the techniques developed in section 9.4. The convergence is uniform in [−1,1].<br />

As usual, the determination of the approximating polynomials is equivalent to<br />

solving a linear system which can be wr<strong>it</strong>ten by defining the vectors<br />

¯pn ≡ pn(η (n)<br />

1 ), · · · ,pn(η (n)<br />

n−1 ) , ¯qn ≡ qn(η (n)<br />

1 ), · · · ,qn(η (n)<br />

n−1 ) ,<br />

¯fn ≡ fn(η (n)<br />

1 ), · · · ,fn(η (n)<br />

n−1 ) , ¯gn ≡ gn(η (n)<br />

1 ), · · · ,gn(η (n)<br />

n−1 ) .


218 Polynomial Approximation of Differential Equations<br />

Then, one has<br />

(9.9.3)<br />

⎡<br />

⎣ Mn<br />

−µIn<br />

⎤⎡<br />

µIn<br />

⎦⎣<br />

¯pn<br />

⎤<br />

⎦ =<br />

Mn<br />

¯qn<br />

⎡<br />

⎣<br />

⎤<br />

¯fn<br />

⎦,<br />

where In denotes the (n − 1) × (n − 1) ident<strong>it</strong>y matrix and Mn is the reduced<br />

(n − 1) × (n − 1) matrix corresponding to problem (7.4.9) (see (7.4.11) for the case<br />

n = 3). In the ultraspherical case w<strong>it</strong>h −1 < α = β ≤ 1, one shows that the matrix<br />

in (9.9.3) has eigenvalues w<strong>it</strong>h pos<strong>it</strong>ive real part (see section 8.2). In add<strong>it</strong>ion, we can<br />

invert the (2n − 2) × (2n − 2) linear system (9.9.3) by blocks, obtaining<br />

(9.9.4)<br />

⎡<br />

⎣ ¯pn<br />

⎤<br />

⎦ =<br />

¯qn<br />

⎡<br />

⎣ (M2 n + µ 2 In) −1 0<br />

0 (M 2 n + µ 2 In) −1<br />

Therefore, we only need to compute the inverse of M 2 n + µ 2 In.<br />

9.10 Integral equations<br />

¯gn<br />

⎤⎡<br />

⎦⎣<br />

Mn<br />

⎤⎡<br />

⎤<br />

−µIn<br />

¯fn<br />

⎦⎣<br />

⎦.<br />

In concluding this chapter, we say a few words about the possibil<strong>it</strong>y of using polynomials<br />

in the approximation of equations in integral form. A standard example is given by the<br />

following singular integral equation:<br />

(9.10.1) A U(x) + B<br />

π<br />

1<br />

−1<br />

U(s)<br />

s − x<br />

µIn<br />

ds = f(x), ∀x ∈] − 1,1[,<br />

where A,B ∈ R are constants and f :] − 1,1[→ R is a given function. The integral<br />

in (9.10.1) is a principal value integral. This is given by the lim<strong>it</strong><br />

(9.10.2) lim<br />

ǫ→0 +<br />

x−ǫ<br />

−1<br />

U(s)<br />

s − x<br />

ds +<br />

1<br />

x+ǫ<br />

Mn<br />

U(s)<br />

s − x ds<br />

<br />

, ∀x ∈] − 1,1[.<br />

¯gn


Ordinary Differential Equations 219<br />

Problem (9.10.1) adm<strong>it</strong>s solutions satisfying appropriate integrabil<strong>it</strong>y cond<strong>it</strong>ions. The-<br />

oretical results have been given in tricomi(1957), kantorovich and krylov(1964),<br />

pogorzelski(1966), hochstadt(1973).<br />

Galerkin and collocation type approximations of the solution U have been studied by<br />

many authors. In the latter approach, the integral is replaced by su<strong>it</strong>able quadrature<br />

sums based on Chebyshev polynomials. Consider the case A = 0 and B = 1. A<br />

classical scheme is obtained by defining V (s) := U(s) √ 1 − s 2 , s ∈] − 1,1[. For n ≥ 1,<br />

recalling (3.4.1) and (3.4.6), we approximate the integral by the formula<br />

(9.10.3)<br />

1<br />

−1<br />

V (s)<br />

s − x<br />

ds<br />

√ 1 − s 2<br />

≈ π<br />

n<br />

n<br />

j=1<br />

V (ξ (n)<br />

j )<br />

ξ (n)<br />

j<br />

x = ξ(n) j , 1 ≤ j ≤ n,<br />

− x,<br />

where the ξ (n)<br />

j ’s are the zeroes of Tn (see (3.1.4)).<br />

Finally, for n ≥ 2, V is approximated by the polynomial pn ∈ Pn−2 satisfying the set<br />

of equations<br />

(9.10.4)<br />

1<br />

n<br />

n<br />

j=1<br />

pn(ξ (n)<br />

j )<br />

ξ (n)<br />

j<br />

− η(n)<br />

i<br />

= f(η (n)<br />

i ), 1 ≤ i ≤ n − 1,<br />

where the η (n)<br />

j ’s are given in (3.1.11).<br />

It is easy to set up the corresponding linear system. Instead, the theoretical analysis of<br />

convergence is not a trivial matter. Hints are provided in krenk(1975). For other results<br />

and generalizations, we quote the following papers: delves(1968), erdogan, gupta<br />

and cook(1973), krenk(1978), elliott(1981), ioakimidis(1981), monegato(1982),<br />

elliott(1982).<br />

* * * * * * * * * * *


220 Polynomial Approximation of Differential Equations<br />

The differential problems examined in this chapter are mainly of academic interest.<br />

Closer to physics applications are the examples of chapter twelve. It is our opinion,<br />

however, that a correct implementation of more realistic problems starts from a full<br />

understanding of the basic patterns.<br />

For simplic<strong>it</strong>y, boundary value problems in two or more variables are not considered.<br />

Actually, although this extension is sometimes straightforward (see chapter thirteen),<br />

most of the times <strong>it</strong> yields a rather involved theoretical analysis, even for simple model<br />

problems, which contrasts w<strong>it</strong>h the elementary expos<strong>it</strong>ion to which we have tried to<br />

adhere.


10<br />

TIME-DEPENDENT PROBLEMS<br />

The purpose of this chapter is to analyze the approximation, using spectral methods,<br />

of differential problems where the solution is time-dependent. We consider parabolic<br />

and hyperbolic partial differential equations in one space variable. Fin<strong>it</strong>e-differences are<br />

usually employed for the treatment of the time variable, hence part of the analysis will<br />

be devoted to this topic.<br />

10.1 The Gronwall inequal<strong>it</strong>y<br />

A basic result, known as Gronwall lemma, will be used in the following. Several versions<br />

exist, but we recall only one of them here.<br />

Theorem 10.1.1 - Let T > 0 and C ≥ 0. Let g and h be two continuous functions<br />

in the interval [0,T], such that<br />

(10.1.1) g(t) ≤ g(0) +<br />

Then, we have<br />

(10.1.2) g(t) ≤ e Ct<br />

t<br />

0<br />

<br />

g(0) +<br />

[Cg(s) + h(s)]ds, ∀t ∈ [0,T].<br />

t<br />

0<br />

h(s)e −Cs <br />

ds , ∀t ∈ [0,T].


222 Polynomial Approximation of Differential Equations<br />

Proof - Using (10.1.1), we can easily check the inequal<strong>it</strong>y G ′ (t) ≤ 0, ∀t ∈ [0,T], where<br />

we defined<br />

G(t) := e −Ct<br />

<br />

g(0) +<br />

t<br />

0<br />

<br />

[Cg(s) + h(s)]ds<br />

−<br />

t<br />

0<br />

h(s)e −Cs ds, t ∈ [0,T].<br />

Therefore, the function G is not increasing. Finally, by (10.1.1), we get<br />

(10.1.3) g(t)e −Ct −<br />

This implies (10.1.2).<br />

t<br />

10.2 Approximation of the heat equation<br />

0<br />

h(s)e −Cs ds ≤ G(t) ≤ G(0) = g(0), ∀t ∈ [0,T].<br />

In the field of linear partial differential equations, the heat equation is a classical example.<br />

The solution U ≡ U(x,t), physically interpreted as a temperature, depends on the space<br />

variable x and the time variable t. In the case of a heated bar I =]−1,1[ of homogeneous<br />

material, the temperature evolves in the time interval [0,T], T > 0, according to the<br />

equation<br />

(10.2.1)<br />

∂U<br />

∂t (x,t) = ζ ∂2U (x,t), ∀x ∈ I, ∀t ∈]0,T],<br />

∂x2 where ζ > 0 is a constant (thermal diffusiv<strong>it</strong>y).<br />

To pose the problem properly, we need an in<strong>it</strong>ial cond<strong>it</strong>ion:<br />

(10.2.2) U(x,0) ≡ U0(x), ∀x ∈ Ī,<br />

where U0 : Ī → R is a given continuous function.


Time-Dependent Problems 223<br />

In add<strong>it</strong>ion, at any t ∈]0,T], su<strong>it</strong>able boundary cond<strong>it</strong>ions at the endpoints of the<br />

interval I are assumed. These are for instance<br />

(10.2.3) U(−1,t) = σ1(t), U(1,t) = σ2(t), ∀t ∈]0,T],<br />

where σ1 : [0,T] → R, σ2 : [0,T] → R, are given continuous functions.<br />

Equation (10.2.1) is of parabolic type. The l<strong>it</strong>erature on this subject is extensive and<br />

cannot be covered here. For a discussion of the general properties of parabolic equa-<br />

tions, we only mention the following comprehensive books: courant and hilbert<br />

(1953), sneddon (1957), greenspan(1961), epstein(1962), sobolev(1964), wein-<br />

berger(1965). The same books can be used as references for the other sections in this<br />

chapter. An approach more closely related to the heat equation is analyzed in widder<br />

(1975), hill and dewynne(1987).<br />

Under appropriate hypotheses on the data U0, σ1 and σ2, existence and unique-<br />

ness of a solution satisfying (10.2.1), (10.2.2), (10.2.3) can be proven. In most cases,<br />

this can be expressed by a su<strong>it</strong>able series expansion.<br />

We note that the compatibil<strong>it</strong>y cond<strong>it</strong>ions<br />

(10.2.4) σ1(0) = U0(−1), σ2(0) = U0(1),<br />

are not required in general. However, though the solution U is very smooth for t ∈]0,T],<br />

if we drop cond<strong>it</strong>ions (10.2.4), heat propagation w<strong>it</strong>h infin<strong>it</strong>e veloc<strong>it</strong>y is manifested at<br />

t = 0. This can adversely affect the numerical analysis. Thus, we assume henceforth<br />

that relations (10.2.4) are satisfied.<br />

The subst<strong>it</strong>ution V (x,t) := U(x,t) − 1<br />

2 (1 − x)σ1(t) − 1<br />

2 (1 + x)σ2(t), x ∈ [−1,1],<br />

t ∈ [0,T], leads to an equivalent formulation w<strong>it</strong>h homogeneous boundary cond<strong>it</strong>ions.<br />

Indeed, we now have<br />

(10.2.5)<br />

∂V<br />

∂t (x,t) = ζ ∂2V (x,t) + f(x,t), ∀x ∈ I, ∀t ∈]0,T],<br />

∂x2 (10.2.6) V (x,0) = U0(x) − 1<br />

2 (1 − x)U0(−1) − 1<br />

2 (1 + x)U0(1), ∀x ∈ Ī,


224 Polynomial Approximation of Differential Equations<br />

(10.2.7) V (−1,t) = V (1,t) = 0, ∀t ∈ [0,T],<br />

where f(x,t) := − 1<br />

2 (1 − x)σ′ 1(t) − 1<br />

2 (1 + x)σ′ 2(t), x ∈ I, t ∈]0,T]. Here, we require<br />

that the functions σ1 and σ2 are differentiable.<br />

We intend to use spectral methods in order to approximate the unknown V . To<br />

this end, for any t ∈ [0,T], the function V (·,t) : I → R is approximated by an<br />

algebraic polynomial pn(·,t) ∈ P 0 n (see (6.4.11)), where the integer n ≥ 2 has been<br />

fixed in advance. For example, we use the collocation method (see section 9.2). The<br />

approximate problem is<br />

(10.2.8)<br />

∂pn<br />

∂t (η(n) i ,t) = ζ ∂2pn ∂x<br />

(η(n)<br />

2 i ,t) + f(η (n)<br />

i ,t), 1 ≤ i ≤ n − 1, ∀t ∈]0,T].<br />

As usual, the nodes η (n)<br />

d (α,β)<br />

i , 1 ≤ i ≤ n − 1, are the zeroes of the polynomial dxP n ,<br />

α > −1, β > −1. In<strong>it</strong>ial and boundary cond<strong>it</strong>ions are respectively given by<br />

(10.2.9) pn(η (n)<br />

i ,0) = U0(η (n)<br />

i ) − 1<br />

2 (1−η(n)<br />

i )U0(−1) − 1<br />

2 (1+η(n)<br />

i )U0(1), 0 ≤ i ≤ n,<br />

(10.2.10) pn(η (n)<br />

0 ,t) = pn(η (n)<br />

n ,t) = 0, ∀t ∈ [0,T].<br />

At this point, our problem is transformed into a (n + 1) × (n + 1) linear system of<br />

ordinary differential equations. Actually, w<strong>it</strong>h the notations of sections 7.2 and 7.4, we<br />

can wr<strong>it</strong>e for any t ∈]0,T] (we choose for instance n = 4):<br />

(10.2.11)<br />

d<br />

dt<br />

⎡<br />

pn(η<br />

⎢<br />

⎣<br />

(n)<br />

1 ,t)<br />

pn(η (n)<br />

⎤ ⎡<br />

⎥ ⎢<br />

⎥ ⎢<br />

2 ,t) ⎥ = ζ ⎢<br />

⎦ ⎣<br />

pn(η (n)<br />

3 ,t)<br />

˜d (2)<br />

11<br />

˜d (2)<br />

21<br />

˜d (2)<br />

31<br />

˜d (2)<br />

12<br />

˜d (2)<br />

22<br />

˜d (2)<br />

32<br />

˜d (2)<br />

13<br />

˜d (2)<br />

23<br />

˜d (2)<br />

33<br />

⎤⎡<br />

pn(η<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎦⎣<br />

(n)<br />

1 ,t)<br />

pn(η (n)<br />

⎤<br />

⎥<br />

2 ,t) ⎥<br />

⎦ +<br />

⎡<br />

f(η<br />

⎢<br />

⎣<br />

(n)<br />

1 ,t)<br />

f(η (n)<br />

⎤<br />

⎥<br />

2 ,t) ⎥<br />

⎦ .<br />

pn(η (n)<br />

3 ,t)<br />

f(η (n)<br />

3 ,t)<br />

Together w<strong>it</strong>h cond<strong>it</strong>ion (10.2.9), this differential system adm<strong>it</strong>s a unique solution.<br />

Moreover, following the remarks of section 8.2, the eigenvalues of the matrix in (10.2.11)<br />

are distinct, real and negative, under su<strong>it</strong>able assumptions on the parameters α and β.


Time-Dependent Problems 225<br />

The next step is to discretize (10.2.11) w<strong>it</strong>h respect to the variable t. This will be<br />

investigated in section 10.6. Let us first analyze the convergence of the solution of the<br />

semi-discrete problem (10.2.8), (10.2.9), (10.2.10) to the solution of problem (10.2.5),<br />

(10.2.6), (10.2.7). Arguing as in section 9.3, we can propose a variational formulation<br />

for equation (10.2.5) as follows:<br />

(10.2.12)<br />

1<br />

−1<br />

1<br />

∂V<br />

∂V ∂(φw)<br />

φw dx = −ζ<br />

∂t −1 ∂x ∂x<br />

dx +<br />

1<br />

−1<br />

fφw dx,<br />

∀φ ∈ X, ∀t ∈]0,T],<br />

where the solution and the test functions both belong to the space X ≡ H 1 0,w(I) (see<br />

(5.7.6)). Existence and uniqueness of weak solutions of (10.2.12) are discussed in lions<br />

and magenes(1972) for the case w ≡ 1.<br />

W<strong>it</strong>h the help of the quadrature formula (3.5.1), the approximating polynomials satisfy<br />

(10.2.13)<br />

n<br />

j=0<br />

+<br />

<br />

∂pn<br />

∂t φ<br />

<br />

(η (n)<br />

j ,t) ˜w (n)<br />

j<br />

n<br />

(fφ)(η (n)<br />

j=0<br />

j ,t) ˜w (n)<br />

1<br />

∂pn ∂(φw)<br />

= −ζ<br />

−1 ∂x ∂x<br />

j , ∀φ ∈ P 0 n, ∀t ∈]0,T].<br />

Interesting properties are obtained from this equation. Consider ν := α = β w<strong>it</strong>h<br />

−1 < ν ≤ 1. For any t ∈]0,T], setting φ(x) := pn(x,t), x ∈ I, one gets<br />

(10.2.14)<br />

=<br />

n<br />

(fpn)(η (n)<br />

j=0<br />

⎛<br />

1 d<br />

⎝<br />

2 dt<br />

n<br />

[pn(η (n)<br />

j=0<br />

j ,t) ˜w (n)<br />

j<br />

1<br />

≤<br />

2<br />

j ,t)] 2 ˜w (n)<br />

j<br />

⎞<br />

n<br />

[pn(η (n)<br />

j=0<br />

⎠ + ζ<br />

1<br />

−1<br />

j ,t)] 2 ˜w (n)<br />

j<br />

∂pn<br />

∂x<br />

1<br />

+<br />

2<br />

∂(pnw)<br />

∂x<br />

dx<br />

dx<br />

n<br />

[f(η (n)<br />

j=0<br />

j ,t)] 2 ˜w (n)<br />

The last inequal<strong>it</strong>y in (10.2.14) is derived from the relation ab ≤ 1<br />

2 (a2 +b 2 ), a,b ∈ R.<br />

j .


226 Polynomial Approximation of Differential Equations<br />

After integration w<strong>it</strong>h respect to the variable t, we have<br />

(10.2.15) pn(·,t) 2 t 1<br />

w,n + 2ζ<br />

0 −1<br />

≤ pn(·,0) 2 w,n +<br />

t<br />

0<br />

∂pn<br />

∂x<br />

∂(pnw)<br />

∂x<br />

<br />

dx ds<br />

<br />

pn(·,s) 2 w,n + Ĩw,nf(·,s) 2 <br />

w,n ds, ∀t ∈]0,T].<br />

The norm ·w,n is defined in (3.8.2) and the interpolation operator Ĩw,n is defined in<br />

section 3.3. Finally, recalling lemma 8.2.1, we can eliminate the integral on the left-hand<br />

side of (10.2.15), since <strong>it</strong> turns out to be pos<strong>it</strong>ive. Therefore, theorem 10.1.1 yields<br />

(10.2.16) pn(·,t) 2 w,n ≤ e t<br />

<br />

pn(·,0) 2 w,n +<br />

t<br />

0<br />

Ĩw,nf(·,s) 2 w,n e −s <br />

ds ,<br />

∀t ∈]0,T].<br />

Taking into account (10.2.9), one shows that the right-hand side in the above expression<br />

is bounded by a constant which does not depend on n. Thus, (10.2.16) shows that the<br />

norm pn(·,t)w (which is equivalent to the norm pn(·,t)w,n by theorem 3.8.2) is<br />

also bounded independently of n. This is interpreted as a stabil<strong>it</strong>y cond<strong>it</strong>ion (compare<br />

w<strong>it</strong>h (9.4.3)). Moreover, using again lemma 8.2.1, we can also give a uniform bound<br />

t ∂pn<br />

to the quant<strong>it</strong>y 0 ∂x (·,s)2wds, ∀t ∈]0,T], by estimating the right-hand side of<br />

(10.2.15) w<strong>it</strong>h the help of (10.2.16).<br />

The analysis of convergence is similar. For any t ∈]0,T], we define χn(t) :=<br />

ˆΠ 1 0,w,nV (·,t) (see (6.4.13) and (9.4.5)). Thus, by (10.2.12) we can wr<strong>it</strong>e<br />

(10.2.17)<br />

+<br />

1<br />

−1<br />

1<br />

∂V<br />

∂χn ∂(φw)<br />

φw dx = −ζ<br />

∂t −1 ∂x ∂x<br />

n<br />

(fφ)(η (n)<br />

j=0<br />

j ,t) ˜w (n)<br />

dx<br />

j , ∀φ ∈ P 0 n, ∀t ∈]0,T].<br />

The sum on the right-hand side of (10.2.17) replaces the integral 1<br />

fφwdx, after<br />

−1<br />

noting that f(·,t) ∈ P1, ∀t ∈]0,T]. Subtracting equation (10.2.13) from equation<br />

(10.2.17) and taking φ := χn − pn, we obtain by lemma 8.2.1


Time-Dependent Problems 227<br />

(10.2.18)<br />

Therefore<br />

(10.2.19)<br />

−<br />

1<br />

1<br />

−1<br />

= −ζ<br />

−1<br />

<br />

∂V<br />

∂t (χn<br />

<br />

− pn) w dx −<br />

1<br />

1<br />

2<br />

−1<br />

∂(χn − pn)<br />

∂x<br />

d<br />

dt χn − pn 2 w,n ≤<br />

∂V<br />

∂t (χn − pn)w dx =<br />

n<br />

j=0<br />

∂[(χn − pn)w]<br />

∂x<br />

n<br />

j=0<br />

<br />

∂pn<br />

∂t (χn<br />

<br />

− pn) (η (n)<br />

j ,t) ˜w (n)<br />

j<br />

dx ≤ 0, ∀t ∈]0,T].<br />

<br />

∂χn<br />

∂t (χn<br />

<br />

− pn) (η (n)<br />

j ,t) ˜w (n)<br />

j<br />

<br />

( ∂χn<br />

∂t ,χn − pn)w,n − ( ∂χn<br />

∂t ,χn<br />

<br />

− pn)w<br />

+ ( ∂(χn − V )<br />

,χn − pn)w, ∀t ∈]0,T].<br />

∂t<br />

The estimate of the error V −pnw ≤ V −χnw+χn−pnw is now a technical matter.<br />

The difference between the inner products (·, ·)w,n and (·, ·)w can be bounded w<strong>it</strong>h<br />

the help of (3.8.15) (see also section 9.4). In add<strong>it</strong>ion, χn converges to V for n → +∞,<br />

by virtue of (9.4.7). Theorem 10.1.1 is used as done in the proof of stabil<strong>it</strong>y. The reader<br />

can easily provide the details.<br />

It turns out that, for any t ∈ [0,T], the error decays for n → +∞, w<strong>it</strong>h a rate only<br />

depending on the regular<strong>it</strong>y of V . However, using inequal<strong>it</strong>y (10.1.2), the multiplicative<br />

term e t appears in the right-hand side of the estimates. Therefore, for large times, the<br />

error bound is not in general very meaningful for practical applications. Refinements<br />

are possible, for example when σ1 ≡ σ2 ≡ 0. In this case, the solution V (·,t) and the<br />

polynomial pn(·,t) are known to decay exponentially to zero for t → +∞.<br />

The convergence analysis of the collocation method for the heat equation is de-<br />

veloped in canuto and quarteroni(1981). The same conclusions hold for the tau<br />

and Galerkin methods. These techniques can be extended to other parabolic operators<br />

or to different boundary cond<strong>it</strong>ions, but very few theoretical results are available. For<br />

homogeneous Neumann boundary cond<strong>it</strong>ions, an analysis of stabil<strong>it</strong>y in the Chebyshev<br />

case is provided in gottlieb, hussaini and orszag(1984). A numerical example for<br />

a nonlinear parabolic equation is presented in section 12.1.


228 Polynomial Approximation of Differential Equations<br />

A special case worth mentioning is I ≡ R in (10.2.1) and (10.2.2). Of course,<br />

we no longer have boundary cond<strong>it</strong>ions. These are supplanted by the cond<strong>it</strong>ion that<br />

U tends to zero at infin<strong>it</strong>y w<strong>it</strong>h a specified minimum rate. We derive a collocation<br />

type approximation in this case. According to escobedo and kavian(1987), first set<br />

V (x,t) := U( √ 4ζxe2ζt ,e4ζt − 1), x ∈ R, t ∈]0, ˆ T], where ˆ T := 1<br />

4ζln(T + 1). Then, <strong>it</strong><br />

is clear that V satisfies the equation<br />

(10.2.20)<br />

∂V<br />

(x,t) = ζ<br />

∂t<br />

2 ∂ V<br />

(x,t) + 2x∂V<br />

∂x2 ∂x (x,t)<br />

<br />

, ∀x ∈ R, ∀t ∈]0, ˆ T],<br />

(10.2.21) V (x,0) = U( 4ζx,0) = U0( 4ζx), ∀x ∈ R.<br />

Assuming that the solution V decays as fast as e −x2<br />

at infin<strong>it</strong>y, we approximate by<br />

Herm<strong>it</strong>e functions (see section 6.7). Let ξ (n)<br />

i , 1 ≤ i ≤ n, be the zeroes of Hn. For any<br />

t ∈]0, ˆ T], we want to find Pn(·,t) ∈ Sn−1 such that<br />

(10.2.22)<br />

∂Pn<br />

∂t (ξ(n)<br />

i ,t) = ζ<br />

2 ∂ Pn<br />

(ξ(n)<br />

∂x2 i ,t) + 2ξ (n)<br />

i<br />

(10.2.23) Pn(ξ (n)<br />

i ,0) = U0( 4ζξ (n)<br />

i ), 1 ≤ i ≤ n.<br />

∂Pn<br />

∂x (ξ(n)<br />

<br />

i ,t) , 1 ≤ i ≤ n,<br />

We can formulate the same problem in the space of polynomials by setting pn := Pne x2<br />

(see section 7.4). This is equivalent to a system of ordinary differential equations. The<br />

corresponding matrix has non-pos<strong>it</strong>ive eigenvalues, as pointed out in section 8.2, which<br />

is very important for proving convergence of Pn to V for n → +∞. Actually,<br />

recalling formula (3.4.1), the proof follows the same arguments as the case when I is<br />

bounded. The theory is presented in funaro and kavian(1988). Finally, we can get<br />

an approximation of the unknown U w<strong>it</strong>h the change of variables: x → x[4ζ(t+1)] −1/2 ,<br />

t → 1<br />

4ζln(t + 1).<br />

We remark that the direct approximation of the unknown U by the Herm<strong>it</strong>e-<br />

collocation method applied to equation (10.2.1) w<strong>it</strong>h I ≡ R, is unstable. This is due<br />

to the fact that the quant<strong>it</strong>y <br />

R ψ′ (ψv) ′ dx is not in general pos<strong>it</strong>ive for ψ ∈ H 1 v(R),<br />

where v(x) := e x2<br />

, x ∈ R.


Time-Dependent Problems 229<br />

We now wr<strong>it</strong>e Pn(x,t) = n−1 −x2<br />

m=0 cm(t)Hm(x)e , where the cm(t)’s are the<br />

Fourier coefficients of Pn at t ∈]0, ˆ T]. Subst<strong>it</strong>uting Pn in (10.2.22) and recalling<br />

(1.7.1), we obtain<br />

(10.2.24)<br />

d<br />

dt cm(t) = −2ζ(m + 1)cm(t), 0 ≤ m ≤ n − 1, ∀t ∈]0, ˆ T].<br />

Now, we can find the explic<strong>it</strong> expression of the coefficients. We note that, from (10.2.23),<br />

the in<strong>it</strong>ial datum is Pn(x,0) = (I∗ v,nU0)( √ 4ζx), v(x) := ex2, x ∈ R (see (6.7.6)).<br />

Another scheme is obtained w<strong>it</strong>h the in<strong>it</strong>ial cond<strong>it</strong>ion Pn(x,0) = (Π ∗ v,nU0)( √ 4ζx),<br />

x ∈ R (see (6.7.3)). This amounts to assigning the values cm(0), 0 ≤ m ≤ n −1, as the<br />

Fourier coefficients of U0. As the reader can easily check, after this modification, the<br />

solution to (10.2.22) turns out to be the approximation of V by the Galerkin method.<br />

Indeed, Pn also satisfies<br />

(10.2.25)<br />

<br />

R<br />

<br />

∂Pn<br />

∂(Pnv) ∂φ<br />

φv dx = −ζ<br />

∂t R ∂x ∂x dx, ∀φ ∈ Sn−1, ∀t ∈]0, ˆ T].<br />

10.3 Approximation of linear first-order problems<br />

We now focus the attention to another time-dependent partial differential equation. The<br />

unknown U : [−1,1] × [0,T] → R, satisfies<br />

(10.3.1)<br />

∂U ∂U<br />

(x,t) = −ζ (x,t), ∀x ∈] − 1,1], ∀t ∈]0,T],<br />

∂t ∂x<br />

(10.3.2) U(x,0) = U0(x), ∀x ∈ [−1,1],<br />

(10.3.3) U(−1,t) = σ(t), ∀t ∈]0,T],


230 Polynomial Approximation of Differential Equations<br />

where ζ > 0, U0 : [−1,1] → R, σ : [0,T] → R, are given. We assume that U0 and<br />

σ are continuous and σ(0) = U0(−1).<br />

All the standard spectral schemes can be considered to approximate U. Neverthe-<br />

less, the theoretical analysis of convergence is more difficult than the analysis of the<br />

problem in section 10.2. We discuss some examples.<br />

Take σ ≡ 0 . In this case, we can explic<strong>it</strong>ly compute the solution, given by U(x,t) =<br />

U0(x − ζt) if − 1 < x − ζt ≤ 1, U(x,t) = 0 if x − ζt ≤ −1. Consider the collocation<br />

method. We must find pn(·,t) ∈ Pn, n ≥ 1, t ∈]0,T], such that<br />

(10.3.4)<br />

∂pn<br />

∂t (η(n)<br />

i ,t) = −ζ ∂pn<br />

∂x (η(n)<br />

i ,t), 1 ≤ i ≤ n, ∀t ∈]0,T],<br />

(10.3.5) pn(η (n)<br />

i ,0) = U0(η (n)<br />

i ), 0 ≤ i ≤ n,<br />

(10.3.6) pn(η (n)<br />

0 ,t) = 0, ∀t ∈]0,T].<br />

Again, these equations can be reduced to a n ×n linear system of ordinary differential<br />

equations. For instance, using the notations of section 7.2, in the case n = 3, we have<br />

(10.3.7)<br />

d<br />

dt<br />

⎡<br />

pn(η<br />

⎢<br />

⎣<br />

(n)<br />

1 ,t)<br />

pn(η (n)<br />

⎤<br />

⎥<br />

2 ,t) ⎥<br />

⎦<br />

pn(η (n)<br />

3 ,t)<br />

⎡<br />

⎢<br />

= −ζ ⎢<br />

⎣<br />

˜d (1)<br />

11<br />

˜d (1)<br />

21<br />

˜d (1)<br />

31<br />

˜d (1)<br />

12<br />

˜d (1)<br />

22<br />

˜d (1)<br />

32<br />

˜d (1)<br />

13<br />

˜d (1)<br />

23<br />

˜d (1)<br />

33<br />

⎤⎡<br />

pn(η<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎦⎣<br />

(n)<br />

1 ,t)<br />

pn(η (n)<br />

⎤<br />

⎥<br />

2 ,t) ⎥<br />

⎥,<br />

t ∈]0,T].<br />

⎦<br />

pn(η (n)<br />

3 ,t)<br />

The eigenvalues of the matrix in (10.3.7) (see also (7.4.3)) have been studied in section<br />

8.1. We will return to this point in section 10.6.<br />

In the Legendre case (w ≡ 1), we can provide a simple analysis of stabil<strong>it</strong>y. In fact,<br />

formula (3.5.1) allows us to wr<strong>it</strong>e for any t ∈]0,T]<br />

(10.3.8)<br />

d<br />

dt pn(·,t) 2 w,n = 2<br />

n<br />

j=0<br />

<br />

∂pn<br />

∂t pn<br />

<br />

(η (n)<br />

j ,t) w (n)<br />

j<br />

=


Time-Dependent Problems 231<br />

= −2ζ<br />

n<br />

j=0<br />

<br />

∂pn<br />

∂x pn<br />

<br />

(η (n)<br />

j ,t) w (n)<br />

j<br />

1<br />

∂pn<br />

= −2ζ<br />

−1 ∂x pn dx = −ζp 2 n(1,t) ≤ 0.<br />

This shows that the norm pn(·,t)w,n, t ∈ [0,T], is bounded by Ĩw,nU0w,n. This<br />

last term is bounded w<strong>it</strong>h respect to n, when U0 is sufficiently regular. A proof of<br />

convergence is given, using similar arguments, by estimating the error pn − Un, where<br />

Un ∈ Pn, n ≥ 1, is a su<strong>it</strong>able projection of the solution U.<br />

Unfortunately, the same techniques are not effective when different weight functions w<br />

are considered. As pointed out in gottlieb and orszag(1977), p.89, in the Chebyshev<br />

case (w(x) = 1/ √ 1 − x 2 , x ∈] − 1,1[), we cannot expect convergence of pn to U in<br />

the norm · w. In the papers of gottlieb and turkel(1985) and salomonoff<br />

and turkel(1989), convergence estimates for Chebyshev-collocation approximations<br />

are proven in a norm weighted by the function w(x) := (1 − x)/(1 + x), x ∈] − 1,1[,<br />

but the theoretical analysis is harder than in the Legendre case.<br />

A different treatment of the boundary cond<strong>it</strong>ions is possible. We modify (10.3.6)<br />

as follows:<br />

(10.3.9)<br />

∂pn<br />

∂t (η(n)<br />

0 ,t) = −ζ ∂pn<br />

∂x (η(n) 0 ,t) − γ pn(η (n)<br />

0 ,t), ∀t ∈]0,T],<br />

where γ > 0. Basically, we are trying to force the polynomial pn to satisfy, at the<br />

point x = −1, both the boundary constraint and the differential equation (the same<br />

trick used in section 9.4 for the Neumann problem). This leads to a (n + 1) × (n + 1)<br />

differential system, where the corresponding matrix is considered in section 7.4 (see<br />

(7.4.6) for the case n = 3). The eigenvalues of this matrix are examined in section<br />

8.1. When γ is proportional to n 2 , a proof of convergence can be given. The Legendre<br />

and Chebyshev cases are examined in funaro and gottlieb(1989) and funaro and<br />

gottlieb(1988) respectively.<br />

Let us now introduce the tau method for σ ≡ 0. This is obtained by projecting<br />

equation (10.3.1) onto the space Pn−1, n ≥ 1. Then, we are concerned w<strong>it</strong>h finding<br />

pn(·,t) ∈ Pn, t ∈ [0,T], such that<br />

<br />

∂pn<br />

(10.3.10) Πw,n−1 (x,t) = −ζ<br />

∂t<br />

∂pn<br />

(x,t), ∀x ∈] − 1,1], ∀t ∈]0,T].<br />

∂x


232 Polynomial Approximation of Differential Equations<br />

According to theorem 6.2.2, we have<br />

(10.3.11)<br />

1<br />

−1<br />

∂pn<br />

∂t<br />

− Πw,n−1<br />

<br />

∂pn<br />

φw dx = 0 =<br />

∂t<br />

d<br />

1<br />

(pn − Πw,n−1pn)φw dx,<br />

dt −1<br />

∀φ ∈ Pn−1, ∀t ∈]0,T],<br />

which shows that the operator ∂/∂t commutes w<strong>it</strong>h the operator Πw,n−1. Therefore,<br />

we can rewr<strong>it</strong>e (10.3.10) as<br />

(10.3.12)<br />

<br />

∂<br />

∂t (Πw,n−1pn)<br />

<br />

(x,t) = −ζ ∂pn<br />

(x,t),<br />

∂x<br />

∀x ∈] − 1,1], ∀t ∈]0,T].<br />

In add<strong>it</strong>ion, we require that pn(−1,t) = 0, ∀t ∈]0,T], and we impose the in<strong>it</strong>ial<br />

cond<strong>it</strong>ion<br />

(10.3.13) pn(·,0) = rΠw,n−1(U0r −1 ), ∀x ∈ [−1,1],<br />

where r(x) := (1 + x), x ∈ [−1,1]. In this way, we also get pn(−1,0) = 0.<br />

Let ck(t), 0 ≤ k ≤ n, t ∈ [0,T], be the Fourier coefficients of pn(·,t) w<strong>it</strong>h respect<br />

to the basis uk ≡ P (α,β)<br />

k , k ∈ N, α > −1, β > −1. Therefore, pn = n k=0 ckuk.<br />

Moreover, by (10.3.12), pn satisfies the set of equations (see section 7.1):<br />

(10.3.14)<br />

d<br />

dt ck(t) = −ζ c (1)<br />

k (t), 0 ≤ k ≤ n − 1, ∀t ∈]0,T].<br />

This is equivalent to a n × n linear system of ordinary differential equations in the<br />

unknowns ck, 0 ≤ k ≤ n−1, where the in<strong>it</strong>ial values ck(0), 0 ≤ k ≤ n−1, are obtained<br />

from (10.3.13) by virtue of the results of section 2.3. We eliminate cn by expressing <strong>it</strong><br />

as a linear combination of the other coefficients w<strong>it</strong>h the help of the relation<br />

(10.3.15) pn(−1,t) =<br />

n<br />

k=0<br />

ck(t)uk(−1) = 0, ∀t ∈ [0,T].<br />

Another characterization is obtained from (10.3.10) by noting that pn satisfies<br />

(10.3.16)<br />

∂pn ∂pn<br />

(x,t) = −ζ<br />

∂t ∂x (x,t) + c′ n(t)un(x), ∀x ∈] − 1,1], ∀t ∈]0,T].<br />

In fact, the last term in (10.3.16) disappears when we apply the projector Πw,n−1.


Time-Dependent Problems 233<br />

Thus, we can finally view pn(·,t) ∈ Pn, w<strong>it</strong>h pn(−1,t) = 0, ∀t ∈ [0,T], as the solution<br />

of the collocation scheme<br />

(10.3.17)<br />

∂pn<br />

∂t (ξ(n)<br />

i ,t) = −ζ ∂pn<br />

∂x (ξ(n)<br />

i ,t), 1 ≤ i ≤ n, ∀t ∈]0,T],<br />

where the ξ (n)<br />

i ’s are the zeroes of un (see section 3.1). By defining qn := pn/r, where<br />

r(x) := (1 + x), x ∈ [−1,1], we have qn(·,t) ∈ Pn−1, ∀t ∈ [0,T]. Using (10.3.17), qn<br />

also satisfies<br />

(10.3.18)<br />

∂qn<br />

∂t (ξ(n)<br />

i ,t) = −ζ<br />

<br />

∂qn<br />

∂x (ξ(n)<br />

(n)<br />

qn(ξ i ,t)<br />

i ,t) +<br />

r(ξ (n)<br />

<br />

, 1 ≤ i ≤ n, ∀t ∈]0,T].<br />

i )<br />

This is equivalent to a n × n linear system of differential equations in the unknowns<br />

qn(ξ (n)<br />

i , ·), 1 ≤ i ≤ n. The entries of the corresponding matrix are (see (7.2.5)):<br />

−ζ[d (1)<br />

ij + δij/(1 + ξ (n)<br />

i )], 1 ≤ i ≤ n, 1 ≤ j ≤ n.<br />

We can prove stabil<strong>it</strong>y when α > −1 and −1 < β ≤ 1. We first note that<br />

1−x<br />

1+x pn = −cnun + {lower degree terms}. Then, (9.3.16) yields<br />

(10.3.19)<br />

= −2ζ<br />

= −2ζ<br />

1<br />

1−x<br />

1+x<br />

−1<br />

1<br />

1−x<br />

1+x<br />

−1<br />

1<br />

d 1−x<br />

dt 1+x<br />

−1<br />

p2nw dx + c2 n(t)un2 <br />

w<br />

1<br />

1−x = 2 1+x<br />

−1<br />

∂pn<br />

∂x pnw dx + 2c ′ n(t)<br />

∂pn<br />

∂x pnw dx = ζ<br />

∂pn<br />

∂t pnw dx + 2c ′ n(t)cn(t)un 2 w<br />

1<br />

−1<br />

1<br />

p<br />

−1<br />

2 n<br />

1−x<br />

1+x unpnw dx + 2c ′ n(t)cn(t)un 2 w<br />

d<br />

<br />

1−x<br />

dx 1+xw <br />

dx ≤ 0, ∀t ∈]0,T].<br />

The last integral is negative by virtue of the restrictions on α and β. This shows that<br />

a certain weighted norm of pn(·,t) is bounded by the in<strong>it</strong>ial data for any t ∈]0,T].<br />

Further details and improvements are given in gottlieb and tadmor(1990).<br />

<br />

A proof of stabil<strong>it</strong>y in a norm which also controls the derivative ∂pn ∂x has been<br />

outlined in gottlieb and orszag(1977), section 8, for the Chebyshev case. The same<br />

proof can be extended to the other Jacobi cases, provided −1 < β ≤ 0 and α > −1.


234 Polynomial Approximation of Differential Equations<br />

Once the stabil<strong>it</strong>y of the scheme is achieved, error estimates are obtained by applying<br />

the same proof to the error pn − Un, where Un ∈ Pn, n ≥ 1 , is a su<strong>it</strong>able projection<br />

of U. The same argument was used in section 10.2 for the heat equation (see also<br />

canuto, hussaini, quarteroni and zang(1988), chapter 10). Other spectral type<br />

approximations and theoretical results relative to equation (10.3.1) are considered in<br />

gottlieb and orszag(1977), section 8, gottlieb(1981), canuto and quarteroni<br />

(1982b), mercier(1982), mercier(1989), tal-ezer(1986b), dubiner(1987).<br />

Similar results are available for the equation<br />

(10.3.20)<br />

∂U ∂U<br />

(x,t) = ζ (x,t), ∀x ∈ [−1,1[, ∀t ∈]0,T],<br />

∂t ∂x<br />

when the boundary cond<strong>it</strong>ion is imposed at the point x = 1.<br />

The reader should pay a l<strong>it</strong>tle more care to the theoretical analysis of partial differ-<br />

ential equations, before trying experiments on general problems, such as the following<br />

one:<br />

(10.3.21)<br />

∂U<br />

(x,t) = A(x,U(x,t))∂U (x,t), x ∈] − 1,1[, t ∈]0,T],<br />

∂t ∂x<br />

where A :] − 1,1[×R → R is given.<br />

An analysis for equations such as (10.3.21) is carried out in smoller(1983), rhee,<br />

aris and amundson(1986), kreiss and lorenz(1989). It is known that the solution<br />

of (10.3.21) maintains a constant value along the so-called characteristic curves. These<br />

are parallel straight-lines in the (x,t) plane w<strong>it</strong>h slope 1/ζ for equation (10.3.1), and<br />

slope −1/ζ for equation (10.3.20). In the former case, the solution U(x, ·), x ∈ [−1,1],<br />

shifts on the right-hand side during the time evolution. W<strong>it</strong>h terminology deriving from<br />

fluid physics, the point x = −1 is the inflow boundary, while x = 1 is the outflow<br />

boundary. The s<strong>it</strong>uation is reversed for equation (10.3.20). The characteristic curves<br />

are no longer straight-lines when A is not a constant. Even in the case of smooth in<strong>it</strong>ial<br />

cond<strong>it</strong>ions and smooth boundary data, the solution can loose regular<strong>it</strong>y and shocks can<br />

be generated when two or more characteristic curves intersect. The numerical analysis<br />

becomes a delicate issue in this s<strong>it</strong>uation, especially for spectral methods which are


Time-Dependent Problems 235<br />

particularly sens<strong>it</strong>ive to the smoothness of the solution. Desp<strong>it</strong>e much effort devoted to<br />

this subject, many problems remain. Polynomial approximations by spectral methods<br />

of equation (10.3.21), when A(x,U) := x or A(x,U) := −x, have been considered in<br />

gottlieb and orszag(1977), section 7. Nonlinear examples will be examined in the<br />

coming section.<br />

10.4 Nonlinear time-dependent problems<br />

An interesting family of nonlinear first order problems is represented by conservation<br />

equations. They are wr<strong>it</strong>ten as<br />

(10.4.1)<br />

∂U<br />

∂t<br />

∂F(U)<br />

(x,t) = (x,t), x ∈] − 1,1[, t ∈]0,T],<br />

∂x<br />

where F : R → R is a given function. If F is differentiable, then (10.4.1) is equivalent to<br />

taking A(x,U) := F ′ (U) in (10.3.21). These equations are consequence of conservation<br />

of energy, mass or momentum, in a physics phenomenon. Boundary cond<strong>it</strong>ions have to<br />

be imposed at the inflow boundaries, which are determined according to the direction<br />

of the characteristic curves. We note that now the slope of a characteristic curve at a<br />

given point (x,t) depends on the unknown U. The case corresponding to F(U) :=<br />

− 1<br />

2 U2 , U ∈ R (hence A(x,U) = −U), is sufficient to illustrate the possible ways in<br />

which the solution can develop. A theoretical discussion is too lengthy to include here.<br />

We suggest the books of smoller(1983), chapter 15, and kreiss and lorenz(1989),<br />

chapter 4, for a more in-depth analysis.<br />

A large variety of fin<strong>it</strong>e-difference schemes has been proposed to solve the above<br />

equations numerically. These are sometimes very sophisticated in order to treat so-<br />

lutions w<strong>it</strong>h sharp derivatives or shocks. Among many celebrated papers, we refer to<br />

sod(1985) for a general introductive overview. There is less published material available<br />

in the field of spectral methods. Algor<strong>it</strong>hms are proposed for the approximation of peri-<br />

odic solutions by trigonometric polynomials, but some of the techniques can be adapted


236 Polynomial Approximation of Differential Equations<br />

to algebraic polynomials. If the solution displays a smooth behavior, a straightforward<br />

application of the usual methods (tau and collocation) generally provides good results.<br />

For the treatment of other s<strong>it</strong>uations, a lot of extra care is needed. Many research groups<br />

are working w<strong>it</strong>h great effort on this subject, following different ideas and devising new<br />

schemes. A detailed analysis of these techniques is an amb<strong>it</strong>ious project that we prefer<br />

to avoid in this book. We will only outline the main patterns. For a closer approach, an<br />

update review and numerous references are given in canuto, hussaini, quarteroni<br />

and zang(1988), chapters 8 and 12.<br />

One strategy commonly used in computations is filtering. The procedure consists in<br />

appropriately damping the higher Fourier coefficients of the approximating polynomial,<br />

to control oscillations near the points of discontinu<strong>it</strong>y of the solution. A drawback to<br />

using filters is that such a smoothing affects the convergence, and reduces the accuracy<br />

of the scheme.<br />

Another approach is to apply domain decompos<strong>it</strong>ion methods (see chapter eleven),<br />

e<strong>it</strong>her to improve the resolution in the cr<strong>it</strong>ical zones, or to allow discontinu<strong>it</strong>ies of the<br />

approximating functions in accordance w<strong>it</strong>h those of the solution. These methods have<br />

been employed for instance in kopriva(1986), macaraeg and streett(1986). The<br />

location of the singular points can be estimated by a procedure called shock-f<strong>it</strong>ting.<br />

Numerical codes based on cell-averaging offer another promising field of research.<br />

Results for Chebyshev approximations are provided in cai, gottlieb and harten<br />

(1990).<br />

As far as the theory is concerned, very l<strong>it</strong>tle is known about these techniques,<br />

especially in the case of algebraic polynomials.<br />

Let us examine some other equations. An interesting variation of (10.4.1) is<br />

(10.4.2)<br />

∂U<br />

(x,t) =<br />

∂t<br />

where ǫ > 0 is a given parameter.<br />

<br />

ǫ ∂2 <br />

U ∂F(U)<br />

+ (x,t), x ∈] − 1,1[, t ∈]0,T],<br />

∂x2 ∂x


Time-Dependent Problems 237<br />

For F(U) := − 1<br />

2 U2 , U ∈ R, (10.4.2) is known as Burgers equation. Due to the presence<br />

of the second-order derivative (which, in analogy w<strong>it</strong>h fluid dynamics, corresponds<br />

to viscos<strong>it</strong>y in the physical model problem), solutions of (10.4.2) are smoother than<br />

solutions corresponding to (10.4.1). This defin<strong>it</strong>ely helps the theoretical analysis. In<br />

add<strong>it</strong>ion, one expects that solutions of (10.4.2) for a small ǫ are in some way close to<br />

solution of (10.4.1). We refer to smoller, p.257, for a theoretical explanation of this<br />

fact. For boundary cond<strong>it</strong>ions of the form U(±1,t) = 0, t ∈]0,T], approximations<br />

of the Burgers equation by Galerkin and collocation methods are respectively consid-<br />

ered in maday and quarteroni(1981), and maday and quarteroni(1982), for the<br />

Chebyshev and Legendre cases. The analysis is carried out for the steady equation,<br />

which means that U does not depend on t (see (9.8.5)). Other results are provided in<br />

bressan and quarteroni(1986a).<br />

Finally, results are also available for the Korteveg-De Vries equation (see pavoni<br />

(1988), bressan and pavoni(1990)):<br />

(10.4.3)<br />

∂U<br />

(x,t) =<br />

∂t<br />

<br />

where ǫ = 0 is a given constant.<br />

ǫ ∂3U ∂U<br />

− U<br />

∂x3 ∂x<br />

10.5 Approximation of the wave equation<br />

<br />

(x,t), x ∈] − 1,1[, t ∈]0,T],<br />

Another classical time-dependent problem is given by the wave equation. Here, the<br />

unknown is the function U : [−1,1] × [0,T] → R satisfying<br />

(10.5.1)<br />

∂2U ∂t2 (x,t) = ζ2 ∂2U (x,t), x ∈] − 1,1[, t ∈]0,T],<br />

∂x2


238 Polynomial Approximation of Differential Equations<br />

(10.5.2) U(±1,t) = 0, ∀t ∈]0,T],<br />

(10.5.3) U(x,0) = U0(x),<br />

∂U<br />

∂t (x,0) = Û0(x), ∀x ∈] − 1,1[,<br />

where ζ > 0 and U0 :] − 1,1[→ R, Û0 :] − 1,1[→ R are given in<strong>it</strong>ial data. Equation<br />

(10.5.1) is of hyperbolic type. The solution U represents the displacement of a vibrat-<br />

ing string of length equal 2, fixed at the endpoints. Other oscillatory phenomena are<br />

described by the wave equation. Examples are collected in baldock and bridgeman<br />

(1981).<br />

From the numerical point of view, we can apply the same techniques of section<br />

10.2 to obtain the semi-discrete spectral approximation of problem (10.5.1), (10.5.2),<br />

(10.5.3). This time, after discretization in the variable x, we get an in<strong>it</strong>ial-value second-<br />

order differential system in the variable t.<br />

In the Legendre case, stabil<strong>it</strong>y results and error estimates can be obtained from a<br />

weak formulation of the problem. The reader is addressed to lions and magenes(1972),<br />

chapter 5, theorem 2.1, for results concerning this aspect of the theory. To our knowl-<br />

edge, no analysis is currently available for the other Jacobi cases.<br />

Another approach consists in wr<strong>it</strong>ing the solution as a su<strong>it</strong>able superpos<strong>it</strong>ion of two<br />

waves travelling in oppos<strong>it</strong>e directions. Actually, one checks that the general solution of<br />

(10.5.1) is given by U(x,t) = τ(x − ζt) + υ(x + ζt), where τ and υ can be uniquely de-<br />

termined by the in<strong>it</strong>ial data (10.5.3) (d’Alembert solution). Thus, U can be decomposed<br />

into the sum of two functions, which are solutions of equations (10.3.1) and (10.3.20)<br />

respectively, w<strong>it</strong>h some prescribed in<strong>it</strong>ial guess. Namely, we have U = V + W, where<br />

V and W solve the following linear symmetric hyperbolic system:<br />

(10.5.4)<br />

⎛ ⎞<br />

V (x,t)<br />

∂<br />

⎝ ⎠ =<br />

∂t<br />

W(x,t)<br />

⎛ ⎞<br />

−ζ 0<br />

⎝ ⎠<br />

0 ζ<br />

∂<br />

⎛ ⎞<br />

V (x,t)<br />

⎝ ⎠ , x ∈] − 1,1[, t ∈]0,T].<br />

∂x<br />

W(x,t)


Time-Dependent Problems 239<br />

Since (10.5.4) is in a diagonal form, V and W are also called characteristic variables.<br />

From (10.5.2), the two unknowns are coupled via the boundary relations<br />

(10.5.5)<br />

where R = L = −1.<br />

⎧<br />

⎨V<br />

(−1,t) = L W(−1,t)<br />

⎩<br />

W(1,t) = R V (1,t)<br />

∀t ∈]0,T],<br />

The physical interpretation of formula (10.5.5) is that the travelling waves are reflected<br />

at the endpoints of the vibrating string. First-order systems, such as (10.5.4), are often<br />

encountered in fluid dynamics models (see kreiss and lorenz(1989), section 7.6).<br />

For spectral type approximations, results are generally obtained for hyperbolic<br />

systems wr<strong>it</strong>ten in terms of characteristic variables which are combined together at the<br />

boundary points by (10.5.5), w<strong>it</strong>h R and L satisfying |RL| < 1 (note that for the wave<br />

equation one has instead |RL| = 1). In this s<strong>it</strong>uation, the ampl<strong>it</strong>ude of the oscillations<br />

decays for t → +∞ (energy dissipation due to friction effects).<br />

The collocation method is obtained in the following way. We are concerned w<strong>it</strong>h<br />

finding pn(·,t) ∈ Pn, qn(·,t) ∈ Pn, n ≥ 1, t ∈]0,T], such that<br />

(10.5.6)<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

∂pn<br />

∂t (η(n)<br />

i ,t) = −ζ ∂pn<br />

∂x (η(n)<br />

i ,t) 1 ≤ i ≤ n, t ∈]0,T],<br />

∂qn<br />

∂t (η(n)<br />

i ,t) = ζ ∂qn<br />

∂x (η(n)<br />

i ,t) 0 ≤ i ≤ n − 1, t ∈]0,T].<br />

At the boundaries, we require that<br />

⎧<br />

⎨pn(−1,t)<br />

= L qn(−1,t)<br />

(10.5.7)<br />

⎩<br />

qn(1,t) = R pn(1,t)<br />

∀t ∈]0,T].<br />

When R and L in (10.5.7) satisfy |LR| < 1, the polynomials pn(·,t) and qn(·,t)<br />

decay to zero when t → +∞ (see lustman(1986)). It is an easy exercise to set up<br />

the 2n × 2n system of ordinary differential equations corresponding to (10.5.6). We<br />

note that two unknowns can be eliminated by virtue of (10.5.7). For n = 2, we get for<br />

example


240 Polynomial Approximation of Differential Equations<br />

(10.5.8)<br />

d<br />

dt<br />

⎡<br />

pn(η<br />

⎢<br />

⎣<br />

(n)<br />

1 ,t)<br />

pn(η (n)<br />

2 ,t)<br />

qn(η (n)<br />

⎤<br />

⎥<br />

0 ,t) ⎥<br />

⎦<br />

qn(η (n)<br />

1 ,t)<br />

⎡<br />

⎢<br />

= ζ ⎢<br />

⎣<br />

− ˜ d (1)<br />

11 − ˜ d (1)<br />

12 −L ˜ d (1)<br />

10<br />

− ˜ d (1)<br />

21 − ˜ d (1)<br />

22 −L ˜ d (1)<br />

20<br />

0 R ˜ d (1)<br />

02<br />

0 R ˜ d (1)<br />

12<br />

˜d (1)<br />

00<br />

˜d (1)<br />

10<br />

0<br />

0<br />

˜d (1)<br />

01<br />

˜d (1)<br />

11<br />

⎤⎡<br />

pn(η<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎦⎣<br />

(n)<br />

1 ,t)<br />

pn(η (n)<br />

2 ,t)<br />

qn(η (n)<br />

⎤<br />

⎥<br />

⎥,<br />

0 ,t) ⎥<br />

⎦<br />

qn(η (n)<br />

1 ,t)<br />

t ∈]0,T].<br />

We note that the determinant of the matrix of the system vanishes when RL = 1.<br />

In fact, polynomials of degree zero satisfying (10.5.7) are eigenfunctions relative to the<br />

eigenvalue zero. Unfortunately, we are not aware of stabil<strong>it</strong>y and convergence results<br />

for this approximation scheme.<br />

Approximations of equation (10.5.4) by polynomials in Pn, using the collocation<br />

method at the points η (n+1)<br />

i , 1 ≤ i ≤ n are considered in gottlieb, lustman<br />

and tadmor(1987a) and gottlieb, lustman and tadmor(1987b). The boundary<br />

cond<strong>it</strong>ions are treated as in (10.5.7), where |RL| < 1. In this case, a theoretical<br />

analysis is given for the ultraspherical case when ν := α = β satisfies −1 < ν ≤ 0.<br />

In funaro and gottlieb(1989), the boundary relations (10.5.7) are modified as<br />

(10.5.9)<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

∂pn<br />

∂pn<br />

(−1,t) = −ζ<br />

∂t ∂x (−1,t) − γ[pn − L qn](−1,t)<br />

∂qn ∂qn<br />

(1,t) = ζ<br />

∂t ∂x (1,t) − γ[qn − R pn](1,t)<br />

∀t ∈]0,T],<br />

where γ > 0 is a constant. As in (10.3.9) the differential equation is also considered at<br />

the points x = ±1. The new collocation scheme is now equivalent to a (2n+2)×(2n+2)<br />

differential system. For n = 2, we have ∀t ∈]0,T]


Time-Dependent Problems 241<br />

(10.5.10)<br />

d<br />

dt<br />

⎡<br />

pn(η<br />

⎢<br />

⎣<br />

(n)<br />

0 ,t)<br />

pn(η (n)<br />

1 ,t)<br />

pn(η (n)<br />

2 ,t)<br />

qn(η (n)<br />

0 ,t)<br />

qn(η (n)<br />

⎤ ⎡<br />

⎥ ⎢<br />

⎥ ⎢<br />

⎥ ⎢<br />

⎥ ⎢<br />

⎥ ⎢<br />

⎥ ⎢<br />

⎥ ⎢<br />

⎥ ⎢<br />

⎥ = ζ ⎢<br />

⎥ ⎢<br />

⎥ ⎢<br />

⎥ ⎢<br />

⎥ ⎢<br />

1 ,t) ⎥ ⎢<br />

⎥ ⎢<br />

⎦ ⎣<br />

qn(η (n)<br />

2 ,t)<br />

− ˜ d (1)<br />

00 − γ − ˜ d (1)<br />

01 − ˜ d (1)<br />

02 γL/ζ 0 0<br />

− ˜ d (1)<br />

10<br />

− ˜ d (1)<br />

20<br />

− ˜ d (1)<br />

11 − ˜ d (1)<br />

12 0 0 0<br />

− ˜ d (1)<br />

21 − ˜ d (1)<br />

22 0 0 0<br />

0 0 0 ˜ d (1)<br />

00<br />

0 0 0 ˜ d (1)<br />

10<br />

0 0 γR/ζ ˜ d (1)<br />

20<br />

˜d (1)<br />

01<br />

˜d (1)<br />

11<br />

˜d (1)<br />

21<br />

˜d (1)<br />

22<br />

˜d (1)<br />

02<br />

˜d (1)<br />

12<br />

− γ<br />

⎤⎡<br />

pn(η<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎦⎣<br />

(n)<br />

0 ,t)<br />

pn(η (n)<br />

1 ,t)<br />

pn(η (n)<br />

2 ,t)<br />

qn(η (n)<br />

0 ,t)<br />

qn(η (n)<br />

⎤<br />

⎥<br />

1 ,t) ⎥<br />

⎦<br />

qn(η (n)<br />

2 ,t)<br />

Convergence estimates for the error |R|pn − V 2 w + |L|qn − W 2 w, are given in<br />

the Legendre case (w ≡ 1) when |RL| < 1 and γ := 1<br />

2 ζn(n + 1) |RL|.<br />

For the proof of stabil<strong>it</strong>y, we argue as follows. Let us assume for simplic<strong>it</strong>y that L > 0<br />

and R > 0 w<strong>it</strong>h LR < 1. By (3.5.6), we first note that ˜w (n)<br />

0<br />

Then, formula (3.5.1) w<strong>it</strong>h w ≡ 1 yields ∀t ∈]0,T]<br />

(10.5.11)<br />

⎛<br />

d<br />

⎝R<br />

dt<br />

n<br />

j=0<br />

1<br />

∂pn<br />

= −2ζR<br />

−1<br />

− 2Rγ ˜w (n)<br />

<br />

0 pn(−1,t) − Lqn(−1,t)<br />

<br />

+<br />

=<br />

<br />

p 2 n(η (n)<br />

j ,t) ˜w (n)<br />

j<br />

∂x pn dx + 2ζL<br />

<br />

+ L<br />

1<br />

−1<br />

n<br />

pn(−1,t) − 2Lγ ˜w (n)<br />

n<br />

q<br />

j=0<br />

2 n(η (n)<br />

j ,t) ˜w (n)<br />

j<br />

= ˜w(n)<br />

n = ζ γ √ RL.<br />

⎞<br />

⎠<br />

∂qn<br />

∂x qn dx<br />

<br />

<br />

qn(1,t) − Rpn(1,t) qn(1,t)<br />

− ζRp 2 n(1,t) + 2γRL ˜w (n)<br />

n pn(1,t)qn(1,t) + L(ζ − 2γ ˜w (n)<br />

n )q 2 <br />

n(1,t)<br />

− ζLq 2 n(−1,t) + 2γRL ˜w (n)<br />

0 pn(−1,t)qn(−1,t) + R(ζ − 2γ ˜w (n)<br />

0 )p 2 <br />

n(−1,t)<br />

=


242 Polynomial Approximation of Differential Equations<br />

<br />

= ζ − Rp 2 n(1,t) + 2 √ RLpn(1,t)qn(1,t) + (L − 2 L/R)q 2 <br />

n(1,t)<br />

<br />

+ ζ − Lq 2 n(−1,t) + 2 √ RLpn(−1,t)qn(−1,t) + (R − 2 R/L)p 2 <br />

n(−1,t)<br />

The last terms in (10.5.11) are negative in view of the inequal<strong>it</strong>ies<br />

2 √ RLab ≤ Ra 2 + Lb 2 ≤ Ra 2 <br />

+ 2 <br />

L/R − L b 2 ,<br />

2 √ RLab ≤ La 2 + Rb 2 ≤ La 2 +<br />

where a,b ∈ R, R > 0, L > 0, RL < 1.<br />

<br />

2 <br />

R/L − R b 2 ,<br />

By (3.8.6), we obtain a bound to the norm Rpn(·,t) 2 w + Lqn(·,t) 2 w, t ∈]0,T], in<br />

terms of the in<strong>it</strong>ial data.<br />

≤ 0.<br />

We also outline the proof of convergence. One introduces two add<strong>it</strong>ional polyno-<br />

mials ˆpn(·,t) ∈ Pn, ˆqn(·,t) ∈ Pn, n ≥ 1, t ∈]0,T]. These are solution of<br />

⎧<br />

(10.5.12)<br />

w<strong>it</strong>h<br />

(10.5.13)<br />

⎪⎨<br />

⎪⎩<br />

∂ˆpn<br />

∂t (η(n)<br />

i ,t) = −ζ ∂ˆpn<br />

∂x (η(n)<br />

i ,t) 1 ≤ i ≤ n, t ∈]0,T],<br />

∂ˆqn<br />

∂t (η(n)<br />

i ,t) = ζ ∂ˆqn<br />

∂x (η(n)<br />

i ,t) 0 ≤ i ≤ n − 1, t ∈]0,T],<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

∂ˆpn<br />

∂ˆpn<br />

(−1,t) = −ζ<br />

∂t ∂x (−1,t) − γ[ˆpn − V ](−1,t)<br />

∂ˆqn ∂ˆqn<br />

(1,t) = ζ<br />

∂t ∂x (1,t) − γ[ˆqn − W](1,t)<br />

∀t ∈]0,T].<br />

The functions ˆpn and ˆqn are decoupled. Therefore, we can estimate the errors<br />

(ˆpn − V )(·,t)w and (ˆqn − W)(·,t)w, t ∈]0,T], by virtue of the results of section<br />

10.3. On the other hand, by subtracting the equations (10.5.12) from the equations<br />

(10.5.6), we get an equivalent collocation scheme in the unknowns rn := pn − ˆpn and<br />

sn := qn − ˆqn , w<strong>it</strong>h the following boundary relations, obtained by combining (10.5.5),<br />

(10.5.9) and (10.5.13):


Time-Dependent Problems 243<br />

(10.5.14)<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

∂rn ∂rn<br />

(−1,t) = −ζ<br />

∂t ∂x (−1,t) − γ[rn − Lsn](−1,t) + γL[ˆqn − W](−1,t)<br />

∂sn ∂sn<br />

(1,t) = ζ<br />

∂t ∂x (1,t) − γ[sn − Rrn](1,t) + γR[ˆpn − V ](1,t)<br />

∀t ∈]0,T].<br />

Adapting the proof of stabil<strong>it</strong>y to this new system, we can bound rn(·,t)w, sn(·,t)w,<br />

t ∈]0,T], by the in<strong>it</strong>ial data and by the errors |ˆpn − V |(1,t), |ˆqn − W |(−1,t), t ∈]0,T].<br />

The final error estimate follows from the triangle inequal<strong>it</strong>ies<br />

pn − V w ≤ rnw + ˆpn − V w, qn − W w ≤ snw + ˆqn − W w.<br />

We finally observe that the numerical experiments are in general more accurate<br />

using (10.5.9) in place of (10.5.7), though the expression of the constant γ is known<br />

only in the Legendre case.<br />

Other polynomial approximations are examined in gottlieb, gurzburger and<br />

turkel(1982) and tal-ezer(1986b). Techniques based on spectral methods are under<br />

development for the numerical discretization of nonlinear hyperbolic systems.<br />

10.6 Time discretization<br />

In the previous sections we analyzed polynomial approximations, w<strong>it</strong>h respect to the<br />

space variable x, of various time-dependent partial differential equations. In order<br />

to implement our algor<strong>it</strong>hms we also need to discretize the time variable t. Although<br />

polynomials can be used to get a global approximation in the time interval [0,T], T > 0<br />

(see morchoisne(1979), tal-ezer(1986a) and tal-ezer(1989)), fin<strong>it</strong>e-differences are<br />

generally preferred. Much has been wr<strong>it</strong>ten on this subject. We refer the reader to


244 Polynomial Approximation of Differential Equations<br />

gear(1971), lapidus and seinfeld(1971), jain(1984). Here, we only present some<br />

basic algor<strong>it</strong>hms.<br />

Our purpose here is to solve first-order differential systems of the form<br />

(10.6.1)<br />

⎧<br />

⎪⎨<br />

d<br />

dt<br />

⎪⎩<br />

¯p(t) = D¯p(t) + ¯ f(t),<br />

¯p(0) = ¯p0,<br />

t ∈]0,T],<br />

where ¯p(t) ∈ R n , t ∈ [0,T], is the unknown. In (10.6.1), D is a n × n matrix,<br />

¯p0 ∈ R n is an in<strong>it</strong>ial guess and ¯ f(t) is a given vector of R n , ∀t ∈]0,T]. Systems like<br />

the one considered above were derived in sections 10.2, 10.3 and 10.5.<br />

The most straightforward time-discretization method is the Euler method. We<br />

subdivide the interval [0,T] in m ≥ 1 equal parts of size h := T/m > 0. Then,<br />

for any m ≥ 1, we construct the vector ¯ph ≡ ¯p (m)<br />

h ∈ Rn , according to the recursion<br />

formula<br />

(10.6.2)<br />

⎧<br />

⎨<br />

⎩<br />

¯p (j)<br />

h := (I + hD)¯p (j−1)<br />

h + h ¯ f(tj−1) 1 ≤ j ≤ m,<br />

¯p (0)<br />

h := ¯p0,<br />

where I denotes the n × n ident<strong>it</strong>y matrix and tj := jh, 0 ≤ j ≤ m. We assume that<br />

D adm<strong>it</strong>s a diagonal form and we denote by λi, 1 ≤ i ≤ n, <strong>it</strong>s eigenvalues. Then, we<br />

have the following result.<br />

Theorem 10.6.1 - Let the eigenvalues of the matrix D satisfy Reλi < 0, 1 ≤ i ≤ n.<br />

Then, if ¯p(t), t ∈ [0,T], is the solution of problem (10.6.1), and ¯ph is obtained by<br />

(10.6.2), <strong>it</strong> is possible to find a constant C > 0 such that<br />

(10.6.3) ¯ph − ¯p(T)R n ≤ Ch sup<br />

for any h satisfying<br />

(10.6.4) 0 < h < ˜ h := min<br />

1≤i≤n<br />

t∈]0,T[<br />

<br />

<br />

<br />

d<br />

<br />

2 <br />

¯p <br />

(t) <br />

dt2 <br />

−2Reλi<br />

|λi| 2<br />

<br />

.<br />

R n<br />

,


Time-Dependent Problems 245<br />

It is well-known that, when cond<strong>it</strong>ion (10.6.4) is violated, severe instabil<strong>it</strong>ies can<br />

occur when applying the algor<strong>it</strong>hm (10.6.2). Actually, (10.6.4) is equivalent to requiring<br />

that the spectral radius ρ(I + hD) is less than 1 (see section 7.6). This stresses the<br />

importance of working w<strong>it</strong>h matrices having eigenvalues w<strong>it</strong>h a negative real part. All<br />

the examples analyzed in the previous sections lead to systems of ordinary differential<br />

equations where the corresponding matrices fulfill such a requirement (see chapter eight).<br />

Moreover, in some cases (see (10.3.8), (10.3.19) and (10.5.11)), we were able to prove an<br />

add<strong>it</strong>ional property, that is, that a certain norm in R n of the vector ¯p(t) is a decreasing<br />

function of t ∈ [0,T]. This automatically excludes the existence of eigenvalues w<strong>it</strong>h a<br />

pos<strong>it</strong>ive real part, otherwise the components of the vector ¯p along the directions of<br />

the corresponding eigenvectors would be amplified during the time evolution. Due to<br />

(10.6.3), the error comm<strong>it</strong>ted at time T decays linearly w<strong>it</strong>h the size of the interval h.<br />

This means that the method is first-order accurate. We observe that the constant C in<br />

(10.6.3) depends on n.<br />

To illustrate the above, we provide the results of a numerical test. Consider equa-<br />

tion (10.2.1) w<strong>it</strong>h ζ = 1 and T = 1. In<strong>it</strong>ial and boundary cond<strong>it</strong>ions are such that<br />

the solution is U(x,t) = e −t cosx + e −4t sinx, x ∈ [−1,1], t ∈ [0,1]. By changing<br />

the variable as suggested in section 10.2, the new unknown V satisfies (10.2.5), (10.2.6)<br />

and (10.2.7). Then, for n ≥ 2, pn denotes the polynomial approximation of V by<br />

the collocation method (10.2.8) at the Chebyshev nodes given in (3.1.11). The relative<br />

differential system (such as (10.2.11) for n = 4) is discretized by the Euler method.<br />

For any m ≥ 1, at the end of the <strong>it</strong>erative process, we obtain a polynomial pn,m ∈ P 0 n<br />

identified by <strong>it</strong>s value at the points η (n)<br />

j , 1 ≤ j ≤ n − 1. In table 10.6.1, we report the<br />

error<br />

En,m :=<br />

1<br />

−1<br />

<br />

pn,m(x) − ( Ĩw,nV )(x,1) <br />

2 dx<br />

√<br />

1 − x2 1<br />

2<br />

, n ≥ 2, m ≥ 1.<br />

From relation (3.8.14), we can evaluate En,m by a quadrature formula based on the<br />

collocation points (see also the numerical examples of section 9.2). The quant<strong>it</strong>y En,m<br />

is the sum of two contributions. One is the error relative to the approximation in<br />

the space variable, which decays exponentially when n grows (compare for instance


246 Polynomial Approximation of Differential Equations<br />

E4,3200 w<strong>it</strong>h E6,3200). The other is the error in the time variable, which only behaves<br />

like h := 1/m. Indeed, except for the case n = 4, En,m is halved when m is doubled.<br />

The reader will note that En,m does not decrease for n ≥ 6. This is because the<br />

time-step h is not small enough to compete w<strong>it</strong>h the good accuracy in space, thus, the<br />

global error is dominated by the time-discretization error. The s<strong>it</strong>uation is different for<br />

the case n = 4, where the space resolution is poor and the results cannot improve by<br />

reducing the time-step. In our example, relation (10.6.4) is not satisfied when n = 10<br />

and m = 200. This is the reason why the error E10,200 is rather high.<br />

m E4,m E6,m E8,m E10,m<br />

200 .5782 × 10 −3 .2426 × 10 −3 .2427 × 10 −3 .3458 × 10 28<br />

400 .5546 × 10 −3 .1212 × 10 −3 .1213 × 10 −3 .1213 × 10 −3<br />

800 .5522 × 10 −3 .6069 × 10 −4 .6066 × 10 −4 .6066 × 10 −4<br />

1600 .5534 × 10 −3 .3055 × 10 −4 .3033 × 10 −4 .3032 × 10 −4<br />

3200 .5545 × 10 −3 .1576 × 10 −4 .1517 × 10 −4 .1516 × 10 −4<br />

Table 10.6.1 - Errors relative to the Chebyshev collocation<br />

approximation of problem (10.2.5), (10.2.6), (10.2.7).<br />

The above experiment shows that, in order to retain the good approximation prop-<br />

erties of spectral methods, the time-step should be very small. This results in a large<br />

number of steps. Although the FFT (see section 4.3) speeds up matrix-vector multipli-<br />

cations in the Chebyshev case, the global computational cost can still be very high. An<br />

alternative is to use more accurate time-discretization methods, such as Runge-Kutta,<br />

Adams-Bashforth or Du Fort-Frankel methods (see jain(1984)). By these techniques,<br />

the error decays as a fixed integer power of the time-step. For instance, experiments<br />

based on equation (10.3.20), using a fourth-order Runge-Kutta method, are discussed in<br />

solomonoff and turkel(1989). Like the Euler method, the schemes mentioned above


Time-Dependent Problems 247<br />

are explic<strong>it</strong>. This implies that they can be used only if h satisfies an inequal<strong>it</strong>y of the<br />

type (10.6.4) (absolute stabil<strong>it</strong>y cond<strong>it</strong>ion). Indications for the choice of the parameter<br />

h are given in gottlieb and tadmor(1990) for spectral approximations of first-order<br />

partial differential equations. Unfortunately, in most of the practical applications, these<br />

stabil<strong>it</strong>y cond<strong>it</strong>ions are considered to be very restrictive, especially when D is related<br />

to the space-discretization of second-order differential operators. In this case, according<br />

to (8.3.6), the maximum time-step ˜ h in (10.6.4) is required to be proportional to 1/n 4 .<br />

Indeed, this lim<strong>it</strong>ation is qu<strong>it</strong>e severe, even when n is not too large. For first-order prob-<br />

lems, kosloff and tal-ezer(1989) propose a Chebyshev-like spectral method w<strong>it</strong>h a<br />

weaker restriction on ˜ h.<br />

No restrictions are in general required for implic<strong>it</strong> methods, such as the backward<br />

Euler method. This is obtained by modifying (10.6.2) to be<br />

(10.6.5)<br />

⎧<br />

⎨<br />

⎩<br />

¯p (j)<br />

h := ¯p (j−1)<br />

h<br />

¯p (0)<br />

h := ¯p0.<br />

+ hD¯p (j)<br />

h + h ¯ f(tj) 1 ≤ j ≤ m,<br />

The scheme (10.6.5) is uncond<strong>it</strong>ionally stable (i.e. no lim<strong>it</strong>ations on h are required),<br />

provided the eigenvalues of D have a negative real part. In add<strong>it</strong>ion, we still have<br />

(10.6.3), i.e., the method is first-order accurate. At each step, the new vector ¯p (j)<br />

h ,<br />

1 ≤ j ≤ m, is computed by solving a n ×n linear system whose matrix is I −hD. We<br />

refer to section 7.6 for the numerical treatment of this system. The algor<strong>it</strong>hm is now<br />

more costly, but we are allowed to choose a larger time-step. This results in a loss of<br />

accuracy, which is moderate, however, for solutions that have a slow time variation.<br />

Further theoretical results and experiments are discussed in mercier(1982) and<br />

mercier(1989) for first-order problems, and canuto and quarteroni(1987) for hy-<br />

perbolic systems. Interesting comments are given in trefethen and trummer(1987)<br />

and trefethen(1988), where numerical tests show the sens<strong>it</strong>iv<strong>it</strong>y to rounding errors<br />

of certain explic<strong>it</strong> methods, when coupled w<strong>it</strong>h spectral approximations.<br />

Similar techniques apply to nonlinear equations. For example, the Burgers equation<br />

(see section 10.4) can be discretized by introducing a sequence of polynomials p (j)<br />

n ∈ P 0 n,


248 Polynomial Approximation of Differential Equations<br />

0 ≤ j ≤ m, according to the recursion formula<br />

(10.6.6) p (j)<br />

n (η (n)<br />

i ) := p (j−1)<br />

n<br />

(η (n)<br />

2 d<br />

i ) + ǫh p(j)<br />

dx2 n (η (n)<br />

<br />

i ) − h<br />

p (j−1)<br />

n<br />

d<br />

dx p(j−1) n<br />

<br />

(η (n)<br />

i ),<br />

1 ≤ i ≤ n − 1, 0 ≤ j ≤ m.<br />

The in<strong>it</strong>ial polynomial is p (0)<br />

n (η (n)<br />

i ) := U0(η (n)<br />

i ), 1 ≤ i ≤ n − 1. The derivatives at<br />

the nodes are evaluated w<strong>it</strong>h the usual arguments (see section 7.2). Then, the final<br />

polynomial p (m)<br />

n<br />

represents an approximation of the exact solution U(·,T). To avoid<br />

severe time-step restrictions the scheme is implic<strong>it</strong>. Since a full implic<strong>it</strong> scheme forces us<br />

to solve a set of nonlinear equations at any step, the nonlinear term is treated explic<strong>it</strong>ly.<br />

A theoretical analysis of convergence of the method for the Chebyshev case is carried<br />

out in bressan and quarteroni(1986a).<br />

The l<strong>it</strong>erature on this subject offers an extensive collection of (more or less effi-<br />

cient) algor<strong>it</strong>hms. An overview of the time-discretization techniques most used in spec-<br />

tral methods can be found in gottlieb and orszag(1977), chapter 9, turkel(1980),<br />

canuto, hussaini, quarteroni and zang(1988), chapter 4, boyd(1989), chapter 8.<br />

A deeper analysis is too technical at this point. Nevertheless, this short introduction<br />

should be sufficient for running the first experiments and interpreting the results.


11<br />

DOMAIN-DECOMPOSITION<br />

METHODS<br />

For several reasons, when approximating differential equations numerically, <strong>it</strong> is often<br />

convenient to decompose the domain, where the solution is defined, into different sub-<br />

sets. Then, independently in each subdomain, one computes polynomial approximations<br />

by the techniques introduced in the previous chapters. The goal now is to find a su<strong>it</strong>able<br />

way to match the different pieces to obtain a discretization of the global solution.<br />

11.1 Introductory remarks<br />

Recently, domain-decompos<strong>it</strong>ion (or multidomain) methods have become a fundamental<br />

subject of research in spectral methods. This is especially true for boundary-value<br />

problems in two or more space variables, when the solutions are defined in domains<br />

w<strong>it</strong>h a complicated geometry. Actually, spectral-type techniques are well su<strong>it</strong>ed for very<br />

simple domains, obtainable by cartesian products of intervals (see chapter thirteen). It<br />

is clear that, when the given domain does not conform w<strong>it</strong>h these requirements, setting<br />

up a spectral approximation scheme is not a straightforward procedure. From this point<br />

of view, this is a severe drawback in comparison w<strong>it</strong>h other more flexible methods, such<br />

as fin<strong>it</strong>e-differences or fin<strong>it</strong>e element methods.


250 Polynomial Approximation of Differential Equations<br />

When possible, the in<strong>it</strong>ial domain is divided into subsets, each one having an ele-<br />

mentary shape. Then, the approximation of the global solution results from combining<br />

the different discrete solutions relative to each subdomain. The way of coupling the<br />

various schemes will be discussed in the coming sections.<br />

In this book we treat only one-dimensional problems. In this context, our domains,<br />

being intervals of R, are simple right from the start. Nevertheless, even for equations<br />

in one variable, there are several s<strong>it</strong>uations where a domain-decompos<strong>it</strong>ion method is<br />

preferable to a direct computation in the whole domain. This is the case for instance<br />

for solutions displaying different regular<strong>it</strong>y behaviors in different parts of the domain.<br />

A global high-degree polynomial approximation is generally more expensive and less<br />

accurate than using high-degree polynomials only in the regions where the degree of<br />

smoothness is lower. In this way, we can treat sharp gradients or shocks when we can<br />

guess their location. Moreover, different techniques can be used in each subdomain<br />

(see the example of section 12.2). The part<strong>it</strong>ioning of the domain is often suggested by<br />

the problem <strong>it</strong>self, as a result of matching different type of equations. In add<strong>it</strong>ion, we<br />

stress that a preliminary theoretical analysis in one variable is often necessary before<br />

experimenting w<strong>it</strong>h more complicated s<strong>it</strong>uations.<br />

We denote by I an open interval of R. Multidomain methods are usually divided<br />

into two categories: non-overlapping and overlapping. In the first case I is decomposed<br />

into a fin<strong>it</strong>e number m ≥ 1 of open intervals Sk, 1 ≤ k ≤ m, in such a way that<br />

Ī = <br />

1≤k≤m ¯ Sk, and Sk1 ∩ Sk2 = ∅, if k1 = k2. In the second case, we assume that<br />

there exist two integers k1 and k2 such that Sk1 ∩ Sk2 = ∅.<br />

11.2 Non-overlapping multidomain methods<br />

Let m ≥ 1 be an integer and sk ∈ R, 0 ≤ k ≤ m, an increasing set of real numbers. We<br />

subdivide the interval I :=]s0,sm[ into the subdomains Sk :=]sk−1,sk[, 1 ≤ k ≤ m.


Domain-Decompos<strong>it</strong>ion Methods 251<br />

Consider first the second-order boundary-value problem (see also (9.1.4)):<br />

(11.2.1)<br />

⎧<br />

⎨ −U ′′ = f in I,<br />

⎩<br />

U(s0) = σ1, U(sm) = σ2,<br />

where σ1,σ2 ∈ R, and f : I → R is a given continuous function.<br />

Given the decompos<strong>it</strong>ion of the domain I, (11.2.1) is equivalent to finding m functions<br />

Uk : ¯ Sk → R, 1 ≤ k ≤ m, such that<br />

(11.2.2)<br />

⎧<br />

−U ′′<br />

k = f in Sk, 1 ≤ k ≤ m,<br />

⎪⎨ Uk(sk) = Uk+1(sk) 1 ≤ k ≤ m − 1,<br />

U ′ k (sk) = U ′ k+1 (sk) 1 ≤ k ≤ m − 1,<br />

⎪⎩<br />

U1(s0) = σ1, Um(sm) = σ2.<br />

Of course, Uk turns out to be the restriction of U to the set ¯ Sk, 1 ≤ k ≤ m.<br />

Instead of approximating the solution of (11.2.1) by a unique polynomial in the whole<br />

interval I as suggested in section 9.4, we look for a continuous approximating function<br />

πn : Ī → R, such that pn,k(x) := πn(x), x ∈ ¯ Sk, is a polynomial in Pn converging<br />

for n → +∞ to the corresponding function Uk, 1 ≤ k ≤ m.<br />

Let us examine for instance the collocation method. First, we map the points<br />

η (n)<br />

j ∈ [−1,1], 0 ≤ j ≤ n, in each interval ¯ Sk, 1 ≤ k ≤ m. This is done by defining the<br />

new set of nodes<br />

(11.2.3) θ (n,k)<br />

j<br />

We note that θ (n,k)<br />

n<br />

:= 1<br />

2 [(sk − sk−1)η (n)<br />

j + sk + sk−1] 0 ≤ j ≤ n, 1 ≤ k ≤ m.<br />

= θ (n,k+1)<br />

0<br />

= sk, 1 ≤ k ≤ m − 1.<br />

Then, following orszag(1980), we define the polynomials pn,k, 1 ≤ k ≤ m, to be<br />

solutions of the set of equations<br />

(11.2.4) −p ′′ n,k(θ (n,k)<br />

i ) = f(θ (n,k)<br />

i ) 1 ≤ i ≤ n − 1, 1 ≤ k ≤ m,


252 Polynomial Approximation of Differential Equations<br />

(11.2.5) pn,k(sk) = pn,k+1(sk) 1 ≤ k ≤ m − 1,<br />

(11.2.6) p ′ n,k(sk) = p ′ n,k+1(sk) 1 ≤ k ≤ m − 1,<br />

(11.2.7) pn,1(s0) = σ1, pn,m(sm) = σ2.<br />

Basically, in (11.2.4) we collocated the differential equation at the nodes inside Sk,<br />

1 ≤ k ≤ m, and imposed the boundary cond<strong>it</strong>ions (11.2.7). Relations (11.2.5) and<br />

(11.2.6) require the continu<strong>it</strong>y of πn and <strong>it</strong>s derivative in the interval I, i.e., πn ∈ C 1 (I).<br />

It is clear that (11.2.4)-(11.2.7) is equivalent to a m(n+1)×m(n+1) linear system. To<br />

determine the corresponding matrix, we argue as in sections 7.2 and 7.4. For instance,<br />

the entries of the derivative matrix at the collocation nodes θ (n,k)<br />

j , 0 ≤ j ≤ n, are given<br />

by 2 ˜ d (1)<br />

ij /(sk − sk−1) 0≤i≤n<br />

0≤j≤n<br />

, 1 ≤ k ≤ m.<br />

In the ultraspherical case, we can replace (11.2.6) by<br />

(11.2.8)<br />

−1<br />

sk+1 − sk−1<br />

+ γk<br />

<br />

(sk − sk−1)p ′′ n,k + (sk+1 − sk)p ′′ <br />

n,k+1 (sk)<br />

<br />

p ′ n,k − p ′ <br />

n,k+1 (sk) = f(sk) 1 ≤ k ≤ m − 1,<br />

where γk, 1 ≤ k ≤ m, are su<strong>it</strong>able constants. In this case we only have πn ∈ C 0 (I).<br />

The use of formula (11.2.8) was proposed in funaro(1986) for Legendre nodes and<br />

funaro(1988) for Chebyshev nodes. This new cond<strong>it</strong>ion, which is closely related to<br />

(9.4.23), results from a variational formulation of problem (11.2.1). When the constants<br />

γk in (11.2.8) are chosen appropriately, numerical results are in general more accurate<br />

than those obtained w<strong>it</strong>h cond<strong>it</strong>ion (11.2.6). In both cases, we expect convergence of<br />

the approximating function πn to the exact solution U when n → +∞. A convergence<br />

analysis can be developed following the variational approach described in sections 9.3<br />

and 9.4. To illustrate the basic principles, we show the theory in a simple case. We<br />

recall that there is no loss of general<strong>it</strong>y if we assume σ1 = σ2 = 0 in (11.2.1).


Domain-Decompos<strong>it</strong>ion Methods 253<br />

Theorem 11.2.1 - Let Uk, 1 ≤ k ≤ m, be the solutions of problem (11.2.2) w<strong>it</strong>h<br />

σ1 = σ2 = 0. Let η (n)<br />

j , 0 ≤ j ≤ n, in (11.2.3), be the nodes of the Legendre (w ≡ 1)<br />

Gauss-Lobatto formula (3.5.1). Let pn,k, 1 ≤ k ≤ m, be the polynomials satisfying<br />

(11.2.4), (11.2.5), (11.2.7) w<strong>it</strong>h σ1 = σ2 = 0, and (11.2.8) w<strong>it</strong>h γk := n(n+1)<br />

sk+1−sk−1 ,<br />

1 ≤ k ≤ m − 1. Then, we have<br />

(11.2.9) lim<br />

n→+∞<br />

sk<br />

sk−1<br />

(Uk − pn,k) 2 dx +<br />

sk<br />

sk−1<br />

Proof - For any 1 ≤ k ≤ m, define the weights ˜w (n,k)<br />

j<br />

(U ′ k − p ′ n,k) 2 dx<br />

1<br />

2<br />

= 0, 1 ≤ k ≤ m.<br />

:= 1<br />

2 (sk −sk−1) ˜w (n)<br />

j , 0 ≤ j ≤ n<br />

(see (3.5.6)). Therefore, by (3.5.1) w<strong>it</strong>h w ≡ 1, we get the integration formula<br />

(11.2.10)<br />

sk<br />

sk−1<br />

p dx =<br />

n<br />

j=0<br />

p(θ (n,k)<br />

j ) ˜w (n,k)<br />

j<br />

∀p ∈ P2n−1, 1 ≤ k ≤ m.<br />

Let Xn denote the space of continuous functions φ in I, such that φ(s0) = φ(sm) = 0<br />

and such that the restriction of φ to any interval ¯ Sk is a polynomial of degree n. Then,<br />

by noting that γk[ ˜w (n,k)<br />

n<br />

satisfies<br />

(11.2.11)<br />

= −<br />

= −<br />

m<br />

<br />

n<br />

k=1<br />

i=1<br />

m<br />

<br />

k=1<br />

= −<br />

Sk<br />

m<br />

k=1<br />

+ ˜w (n,k+1)<br />

0 ] = 1, 1 ≤ k ≤ m − 1, the function πn ∈ Xn<br />

<br />

I<br />

π ′′<br />

nφ dx +<br />

sk<br />

sk−1<br />

[p ′′ n,kφ](θ (n,k)<br />

i ) ˜w (n,k)<br />

i<br />

=<br />

m<br />

<br />

n<br />

k=1<br />

i=1<br />

π ′ nφ ′ dx =<br />

m−1 <br />

k=1<br />

<br />

p ′′ n,kφ dx +<br />

<br />

+<br />

m−1 <br />

k=1<br />

[fφ](θ (n,k)<br />

i ) ˜w (n,k)<br />

i<br />

m<br />

<br />

k=1<br />

lim<br />

x→s −<br />

k<br />

m−1 <br />

k=1<br />

Sk<br />

π ′ nφ ′ dx<br />

(π ′ nφ)(x) − lim<br />

x→s +<br />

k<br />

(π ′ <br />

nφ)(x)<br />

<br />

p ′ n,kφ − p ′ <br />

n,k+1φ (sk)<br />

γk( ˜w (n,k)<br />

n<br />

<br />

+ ˜w (n,k+1)<br />

<br />

0 ) p ′ n,kφ − p ′ <br />

n,k+1φ (sk)<br />

=: Fn(φ), ∀φ ∈ Xn.


254 Polynomial Approximation of Differential Equations<br />

Thus, our collocation scheme is now in a variational form. On the other hand, U satisfies<br />

<br />

I U ′ ψ ′ dx = <br />

I fψdx, ∀ψ ∈ X, ψ(s0) = ψ(sm) = 0, where X is the Hilbert space<br />

w<strong>it</strong>h norm ψX = <br />

I ψ2dx + <br />

I [ψ′ ] 2dx 1/2 . Due to theorem 9.4.1, <strong>it</strong> is sufficient to<br />

estimate the quant<strong>it</strong>y <br />

I fφdx − Fn(φ) , ∀φ ∈ Xn, to conclude that the error |U −πn|<br />

tends to zero in the norm of X. To this end, for any Sk, we can argue as in (9.4.12),<br />

(9.4.13), which lead to (11.2.9).<br />

W<strong>it</strong>h some technical modifications, we can use the same proof to obtain a conver-<br />

gence result for the scheme (11.2.4), (11.2.5), (11.2.6), (11.2.7), in the Legendre case.<br />

For other sets of collocation nodes the theory becomes more difficult. The idea is to<br />

recast the collocation scheme into a variational framework which then reduces to a<br />

problem like (9.4.16) in each subdomain.<br />

We get similar conclusions when different degrees nk, 1 ≤ k ≤ m, are assigned to<br />

the approximating polynomials. Now, we can e<strong>it</strong>her use cond<strong>it</strong>ion (11.2.6) as is, or con-<br />

d<strong>it</strong>ion (11.2.8) after su<strong>it</strong>able modification. For example, in the Legendre case, starting<br />

from the variational formulation (11.2.11), we deduce that the following relations are<br />

satisfied at the interface nodes:<br />

<br />

(11.2.12) − ˜w (nk,k)<br />

nk<br />

+ γk<br />

<br />

where γk := 1 ( ˜w (nk,k)<br />

nk<br />

+ ˜w (nk+1,k+1)<br />

−1 <br />

0<br />

˜w (nk,k)<br />

nk<br />

p ′′ nk,k + ˜w (nk+1,k+1)<br />

0 p ′′ <br />

nk+1,k+1 (sk)<br />

<br />

p ′ nk,k − p ′ <br />

nk+1,k+1 (sk) = f(sk) 1 ≤ k ≤ m − 1,<br />

+ ˜w (nk+1,k+1)<br />

0 ), 1 ≤ k ≤ m − 1.<br />

In the applications, we can prescribe the polynomial degrees nk, depending on the<br />

regular<strong>it</strong>y of U in the different subintervals Sk. We show an example for I =] − 1,1[.<br />

In (11.2.1) we take σ1 = σ2 = 0 and f such that U(x) := sin<br />

12π<br />

5x+7<br />

<br />

, x ∈ Ī.<br />

The function U is plotted in figure 11.2.1. In the first test, we approximate U by a<br />

single polynomial pn w<strong>it</strong>h the standard collocation method at the Legendre points<br />

(see (9.2.15)). In table 11.2.1, we give the error En :=<br />

for different values of n.<br />

n j=1 (pn − U) 2 (η (n)<br />

j ) ˜w (n)<br />

j<br />

1<br />

2


Domain-Decompos<strong>it</strong>ion Methods 255<br />

n En<br />

12 .557106<br />

16 .4418 × 10 −1<br />

20 .4262 × 10 −2<br />

24 .6324 × 10 −3<br />

28 .4265 × 10 −4<br />

32 .7923 × 10 −5<br />

Figure 11.2.1 - Plot of the Table 11.2.1 - Errors for a single-domain<br />

function U(x), x ∈ [−1,1]. approximation of problem (11.2.1).<br />

In the second experiment, we consider two domains, namely S1 =]s0,s1[=] − 1,0[ and<br />

S2 =]s1,s2[=]0,1[. In each interval Sk, 1 ≤ k ≤ 2, we approximate the solution by a<br />

polynomial pnk,k ∈ Pnk , w<strong>it</strong>h the collocation method at the Legendre nodes. At the<br />

point s1 = 0 we impose pn1,1(s1) = pn2,2(s1) and cond<strong>it</strong>ion (11.2.12) for m = 2. We<br />

report in table 11.2.2 the error En1,n2 :=<br />

2<br />

k=1<br />

for various choices of the parameters n1 and n2.<br />

nk<br />

j=1 (pnk,k − U) 2 (θ (nk,k)<br />

j<br />

n1 En1,2 En1,4 En1,6 En1,8<br />

) ˜w (nk,k)<br />

j<br />

12 .130794 .1817 × 10 −1 .1839 × 10 −1 .1839 × 10 −1<br />

16 .133438 .1645 × 10 −2 .1022 × 10 −2 .1021 × 10 −2<br />

20 .133418 .1274 × 10 −2 .5449 × 10 −4 .2638 × 10 −4<br />

24 .133418 .1274 × 10 −2 .4772 × 10 −4 .2089 × 10 −5<br />

28 .133418 .1274 × 10 −2 .4769 × 10 −4 .9319 × 10 −6<br />

Table 11.2.2 - Errors for a multidomain approximation of problem (11.2.1).<br />

1<br />

2


256 Polynomial Approximation of Differential Equations<br />

Due to the different behavior of U in the two subdomains, the best results are<br />

obtained w<strong>it</strong>h an uneven distribution of nodes (n1 > n2). A correct balance of the two<br />

parameters guarantees performances as good as the ones of the single-domain approxi-<br />

mation, w<strong>it</strong>h the same degrees of freedom and a lower computational cost. In fact, the<br />

matrices involved in the multidomain approach are block-diagonal, w<strong>it</strong>h a number of<br />

blocks equal to the number of subdomains and a bandwidth proportional to the largest<br />

degree used. We can take advantage of this structure for a more efficient numerical<br />

implementation. The details are discussed in section 11.3. In (11.2.23), we explic<strong>it</strong>ly<br />

wr<strong>it</strong>e the system from the above example w<strong>it</strong>h n1 = n2 = 2 (hence γ1 = 3):<br />

(11.2.13)<br />

⎡<br />

⎢<br />

⎣<br />

1 0 0 0 0 0<br />

−4 ˜ d (2)<br />

10 −4 ˜ d (2)<br />

11 −4 ˜ d (2)<br />

12 0 0 0<br />

0 0 1 −1 0 0<br />

τ1 τ2 τ3 τ4 τ5 τ6<br />

0 0 0 −4 ˜ d (2)<br />

10 −4 ˜ d (2)<br />

11 −4 ˜ d (2)<br />

12<br />

0 0 0 0 0 1<br />

⎤⎡<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎥⎢<br />

⎦⎢<br />

⎣<br />

pn1,1(θ (n1,1)<br />

0<br />

pn1,1(θ (n1,1)<br />

1<br />

pn1,1(θ (n1,1)<br />

2<br />

pn2,2(θ (n2,2)<br />

0<br />

pn2,2(θ (n2,2)<br />

1<br />

pn2,2(θ (n2,2)<br />

2 )<br />

⎤ ⎡<br />

)<br />

⎥ ⎢<br />

) ⎥ ⎢<br />

⎥ ⎢<br />

⎥ ⎢<br />

⎥ ⎢<br />

) ⎥ ⎢<br />

⎥ ⎢<br />

⎥ ⎢<br />

⎥=<br />

⎢<br />

) ⎥ ⎢<br />

⎥ ⎢<br />

⎥ ⎢<br />

⎥ ⎢<br />

) ⎥ ⎢<br />

⎥ ⎢<br />

⎦ ⎣<br />

σ1<br />

⎥<br />

) ⎥<br />

0 ⎥<br />

f(s1) ⎥<br />

) ⎥<br />

⎦<br />

f(θ (n1,1)<br />

1<br />

f(θ (n2,2)<br />

1<br />

where τ1 := −2 ˜ d (2)<br />

20 + 2γ1 ˜ d (1)<br />

20 , τ2 := −2 ˜ d (2)<br />

21 + 2γ1 ˜ d (1)<br />

21 , τ3 := −2 ˜ d (2)<br />

22 + 2γ1 ˜ d (1)<br />

22 ,<br />

τ4 := −2 ˜ d (2)<br />

00 −2γ1 ˜ d (1)<br />

00 , τ5 := −2 ˜ d (2)<br />

01 −2γ1 ˜ d (1)<br />

01 , τ6 := −2 ˜ d (2)<br />

02 −2γ1 ˜ d (1)<br />

02 . We can remove<br />

one unknown by eliminating the third row and summing up the third and the fourth<br />

columns. Similarly, cond<strong>it</strong>ion (11.2.6) is imposed by defining τ1 := 2 ˜ d (1)<br />

20 , τ2 := 2 ˜ d (1)<br />

21 ,<br />

τ3 := 2 ˜ d (1)<br />

22 , τ4 := −2 ˜ d (1)<br />

00 , τ5 := −2 ˜ d (1)<br />

01 , τ6 := −2 ˜ d (1)<br />

02 , and by setting to zero the<br />

fourth entry of the right-hand side vector.<br />

In the Chebyshev case, relation (11.2.6) is more su<strong>it</strong>able for implementation since<br />

<strong>it</strong> is difficult to obtain the exact expression of γk in (11.2.8), or the counterpart of<br />

formula (11.2.12). As an alternative, we can use the cond<strong>it</strong>ion<br />

σ2<br />


Domain-Decompos<strong>it</strong>ion Methods 257<br />

(11.2.14) p ′ nk,k(sk−1) − p ′ nk+1,k+1(sk+1) =<br />

which is suggested by the ident<strong>it</strong>y<br />

(11.2.15) U ′ (sk−1) − U ′ sk+1<br />

(sk+1) = −<br />

sk−1<br />

sk+1<br />

sk−1<br />

U ′′ dx =<br />

f dx 1 ≤ k ≤ m − 1,<br />

sk+1<br />

sk−1<br />

f dx 1 ≤ k ≤ m − 1.<br />

The integral in (11.2.14) can be computed by quadrature, for instance by the Clenshaw-<br />

Curtis formula once the values of f at the nodes are known (see section 3.7). This<br />

approach was suggested in macaraeg and streett(1986) and produces results com-<br />

parable to those based on relation (11.2.6). Going back to the example previously<br />

illustrated, (11.2.14) w<strong>it</strong>h m = 2 is obtained by setting in (11.2.13): τ1 := 2 ˜ d (1)<br />

00 ,<br />

τ2 := 2 ˜ d (1)<br />

01 , τ3 := 2 ˜ d (1)<br />

02 , τ4 := −2 ˜ d (1)<br />

20 , τ5 := −2 ˜ d (1)<br />

21 , τ6 := −2 ˜ d (1)<br />

22 , and by replacing<br />

the fourth entry of the right-hand side vector by 1<br />

−1 fdx.<br />

The spectral element method (see patera(1984) and section 9.6) uses in the Cheby-<br />

shev case a variational formulation similar to that in (11.2.11) (we recall that this cor-<br />

responds to an approximation of problem (11.2.1) w<strong>it</strong>h σ1 = σ2 = 0). In any interval<br />

¯Sk, 1 ≤ k ≤ m, we consider the expansion pn,k = n<br />

j=0<br />

(11.2.16) ˜ l (n,k)<br />

j<br />

pn,k(θ (n,k)<br />

j<br />

) ˜l (n,k)<br />

j , where<br />

(x) := ˜l (n) <br />

j (2x − sk − sk−1)/(sk − sk−1) , ∀x ∈ ¯ Sk, 0 ≤ j ≤ n.<br />

The nodes are obtained from (11.2.3), where η (n)<br />

j , 0 ≤ j ≤ n, are the Gauss-Lobatto<br />

Chebyshev points (see (3.1.11)). The associated Lagrange polynomials ˜ l (n)<br />

j , 0 ≤ j ≤ n,<br />

are defined in (3.2.10). Next, we modify the right-hand side of (11.2.11) by defining<br />

(11.2.17)<br />

m<br />

<br />

n<br />

Fn(φ) := f(θ (n,k)<br />

<br />

i ) ˜(n,k) l i φ dx ,<br />

Sk<br />

∀φ ∈ Xn.<br />

k=1<br />

i=1<br />

The set of equations is obtained by testing on the functions φ ≡ φk,j ∈ Xn, such that<br />

⎧<br />

⎨˜(n,k)<br />

l j (x)<br />

φk,j(x) :=<br />

⎩<br />

if x ∈ ¯ 0<br />

Sk<br />

if x ∈ Ī − ¯ Sk<br />

1 ≤ j ≤ n − 1, 1 ≤ k ≤ m,


258 Polynomial Approximation of Differential Equations<br />

and on the set of functions φ ≡ φk ∈ Xn, such that<br />

⎧<br />

˜ l (n,k)<br />

n (x) if x ∈ ¯ Sk<br />

⎪⎨<br />

φk(x) :=<br />

⎪⎩<br />

0 elsewhere<br />

˜ l (n,k+1)<br />

0 (x) if x ∈ ¯ Sk+1<br />

1 ≤ k ≤ m − 1.<br />

The cumbersome construction of the corresponding (nm−1)×(nm−1) linear system is<br />

left to the reader. Since we are using Chebyshev nodes instead of Legendre nodes, we can<br />

apply the FFT (see section 4.3) in our computations. Besides, the proof of convergence<br />

follows from theorem 9.4.1 and by showing that the term <br />

I fφdx − Fn(φ) , where<br />

φ ∈ Xn, tends to zero. Different polynomial degrees may be also considered in each<br />

domain.<br />

We can use the same techniques introduced above when the approximating poly-<br />

nomials are represented in the frequency space. For example, the multidomain tau<br />

method is obtained by seeking pn,k = n j=0 cj,kuj,k, 1 ≤ k ≤ m, where uj,k(x) :=<br />

<br />

(2x − sk − sk−1)/(sk − sk−1) , x ∈ ¯ Sk, 1 ≤ k ≤ m, j ∈ N, and the uj’s are the<br />

uj<br />

classical orthogonal polynomials in [−1,1]. Then, in analogy w<strong>it</strong>h (11.2.4)-(11.2.7), we<br />

require that<br />

(11.2.18) − c (2)<br />

j,k<br />

(11.2.19)<br />

(11.2.20)<br />

(11.2.21)<br />

n<br />

j=0<br />

n<br />

j=0<br />

cj,k uj,k(sk) =<br />

cj,k u ′ j,k(sk) =<br />

n<br />

j=0<br />

= fj,k<br />

1<br />

n<br />

j=0<br />

n<br />

j=0<br />

cj,1 uj,1(s0) = σ1,<br />

≤ k ≤ m,<br />

cj,k+1 uj,k+1(sk) 1 ≤ k ≤ m − 1,<br />

cj,k+1 u ′ j,k+1(sk) 1 ≤ k ≤ m − 1,<br />

n<br />

cj,m uj,m(sm) = σ2.<br />

j=0


Domain-Decompos<strong>it</strong>ion Methods 259<br />

In (11.2.18), the c (2)<br />

j,k ’s are the Fourier coefficients of the polynomial p′′ n,k , which,<br />

thanks to the results of section 7.1, can be expressed in terms of the cj,k’s. For any<br />

1 ≤ k ≤ m, the quant<strong>it</strong>ies fj,k, j ∈ N, are the Fourier coefficients of f in ¯ Sk w<strong>it</strong>h<br />

respect to uj,k, j ∈ N. In the Legendre case, (11.2.18)-(11.2.21) are equivalent to the<br />

variational problem (11.2.11), for a su<strong>it</strong>able right-hand side function F.<br />

We finally note that the Legendre-Galerkin method is obtained by modifying the<br />

right-hand side of (11.2.11) by F(φ) := <br />

fφdx, ∀φ ∈ Xn.<br />

I<br />

Of course, different discretizations can be used in each subdomain. An example<br />

of a problem defined in an unbounded domain, solved by coupling Legendre and La-<br />

guerre approximations, is presented in section 12.2. The coupling of fin<strong>it</strong>e element and<br />

spectral methods is considered in bernardi, maday and sacchi-landriani(1989) and<br />

bernardi, deb<strong>it</strong> and maday(1990).<br />

First-order problems are approached in a similar way. For example, the differential<br />

equation (see also (9.1.3))<br />

(11.2.22)<br />

can be wr<strong>it</strong>ten as<br />

(11.2.23)<br />

⎧<br />

⎪⎨<br />

⎧<br />

⎨U<br />

′ + AU = f in ]s0,sm],<br />

⎩<br />

U(s0) = σ,<br />

U ′ k + AUk = f in ]sk−1,sk[, 1 ≤ k ≤ m − 1,<br />

U ′ m + AUm = f in ]sm−1,sm],<br />

Uk(sk) = Uk+1(sk) 1 ≤ k ≤ m − 1,<br />

⎪⎩<br />

U1(s0) = σ,<br />

where Uk is the restriction of U to the set [sk−1,sk], 1 ≤ k ≤ m.<br />

A multidomain collocation scheme is obtained by finding the polynomials pn,k ∈ Pn,<br />

1 ≤ k ≤ m, such that<br />

(11.2.24) (p ′ n,k + Apn,k)(θ (n,k)<br />

i ) = f(θ (n,k)<br />

i ) 1 ≤ i ≤ n − 1, 1 ≤ k ≤ m − 1,<br />

(11.2.25) (p ′ n,m + Apn,m)(θ (n,m)<br />

i ) = f(θ (n,m)<br />

i ) 1 ≤ i ≤ n,


260 Polynomial Approximation of Differential Equations<br />

(11.2.26) pn,k(sk) = pn,k+1(sk) 1 ≤ k ≤ m − 1,<br />

(11.2.27) pn,1(s0) = σ.<br />

Another approximation is obtained by replacing cond<strong>it</strong>ion (11.2.26) by<br />

(11.2.28)<br />

1<br />

sk+1 − sk−1<br />

<br />

(sk −sk−1)(p ′ n,k+Apn,k) + (sk+1 −sk)(p ′ <br />

n,k+1+Apn,k+1) (sk)<br />

<br />

+ γk pn,k − pn,k+1 (sk) = f(sk) 1 ≤ k ≤ m − 1,<br />

where γk, 1 ≤ k ≤ m, are su<strong>it</strong>able constants. We note that using (11.2.28) the global<br />

approximating function in [s0,sm] is not continuous. Nevertheless, we still observe a<br />

spectral rate of convergence.<br />

Cond<strong>it</strong>ions at the interface points for linear or nonlinear first-order time-dependent<br />

problems (see sections 10.3, 10.4 and 10.5) are considered for instance in patera(1984),<br />

kopriva(1986), macaraeg and streett(1986), canuto and quarteroni(1987), fu-<br />

naro(1990b). Other time-dependent equations are examined in pavoni(1988) and<br />

bressan and pavoni(1990). These techniques are under development and very l<strong>it</strong>tle is<br />

known about the theory. Finally, the matching of equations having different orders in<br />

the different domains is examined in gastaldi and quarteroni(1989).<br />

11.3 Solution techniques<br />

The matrices related to the numerical approximation of a differential problem by domain-<br />

decompos<strong>it</strong>ion can be part<strong>it</strong>ioned into diagonal blocks, coupled through the cond<strong>it</strong>ions<br />

imposed at the interface points. The number of blocks and their size is arb<strong>it</strong>rary. Some<br />

authors are inclined to use many subdomains and low polynomial degrees. As in the<br />

fin<strong>it</strong>e element method, they achieve convergence to the exact solution by letting the


Domain-Decompos<strong>it</strong>ion Methods 261<br />

size of these domains tend to zero. This is the case of the h-p-version of the fin<strong>it</strong>e<br />

element method (see babu˘ska, szabo and katz(1981), babu˘ska and suri(1987)). In<br />

contrast, the common practice in spectral methods is to use large domains and high<br />

polynomial degrees. Whether the first approach is better than the latter depends on<br />

the circumstances.<br />

The use of domain-decompos<strong>it</strong>ion methods gives us the possibil<strong>it</strong>y of developing<br />

useful <strong>it</strong>erative algor<strong>it</strong>hms by taking advantage of the structure of the matrices. At each<br />

<strong>it</strong>eration, the linear equations in the system are decoupled by blocks, which are solved<br />

independently. Of course, such a feature is particularly well adapted to computers w<strong>it</strong>h<br />

a parallel arch<strong>it</strong>ecture. We illustrate the basic ideas w<strong>it</strong>h an example. We study the<br />

approximation πn of problem (11.2.1), given by the multidomain collocation scheme<br />

(11.2.4)-(11.2.7). We first introduce some notations. Let yk, 0 ≤ k ≤ m, be a set of<br />

real values w<strong>it</strong>h y0 := σ1 and ym := σ2. For any 1 ≤ k ≤ m, we define the polynomial<br />

qn,k ∈ Pn satisfying the set of equations<br />

(11.3.1)<br />

⎧<br />

⎨ −q ′′<br />

n,k (θ(n,k) i ) = f(θ (n,k)<br />

i ) 1 ≤ i ≤ n − 1,<br />

⎩<br />

qn,k(sk−1) = yk−1, qn,k(sk) = yk.<br />

Next, we define the mapping Γ : R m−1 → R m−1 , Γ ≡ (Γ1, · · · ,Γm−1), whose compo-<br />

nents are given by<br />

(11.3.2) Γk(y1, · · · ,ym−1) := q ′ n,k(sk) − q ′ n,k+1(sk), 1 ≤ k ≤ m − 1.<br />

It is clear that πn is characterized by the relation<br />

(11.3.3) Γk(πn(s1), · · · ,πn(sm−1)) = 0, 1 ≤ k ≤ m − 1.<br />

From the values πn(sk), 1 ≤ k ≤ m − 1, <strong>it</strong> is easy to recover the entire function πn<br />

by noting that qn,k in (11.3.1) coincides w<strong>it</strong>h pn,k when yk−1 = πn(sk−1) and<br />

yk = πn(sk). The problem is consequently reduced to finding the m − 1 zeroes of<br />

Γ. This can be done by an <strong>it</strong>erative procedure. Using an explic<strong>it</strong> method, at each<br />

step we solve in parallel m − 1 linear systems of dimension (n + 1) × (n + 1) corre-<br />

sponding to (11.3.1). This suggests to constructing once for all, for any 1 ≤ k ≤ m,


262 Polynomial Approximation of Differential Equations<br />

an operator, that for any pair of input values (yk−1,yk) in (11.3.1) provides the out-<br />

put values (q ′ n,k (sk−1),q ′ n,k (sk)). For f ≡ 0, σ1 = σ2 = 0, the mapping Γ is<br />

linear and corresponds to the Schur complement of the matrix associated w<strong>it</strong>h (11.2.4)-<br />

(11.2.7), w<strong>it</strong>h respect to the unknowns corresponding to the interface points (see duff,<br />

erisman and reid(1986), p.60). Several <strong>it</strong>erative techniques are available for the nu-<br />

merical determination of the zeroes of Γ. Convergence results in an abstract context<br />

are given in agoshkov(1988). The conjugate gradient method, which is theoretically<br />

examined in the case of fin<strong>it</strong>e element multidomain approximations in bjørstad and<br />

widlund(1986), is studied in chan and goovaerts(1989) for spectral Legendre ap-<br />

proximations. We note that the rate of convergence of the method depends on the ratio<br />

between the maximum and the minimum size of the domains.<br />

Variants of the algor<strong>it</strong>hm proposed above have been analyzed recently. An <strong>it</strong>era-<br />

tive approach is obtained by alternating Dirichlet and Neumann boundary cond<strong>it</strong>ions<br />

in the solution of the problems in the different subdomains. The solution in a given<br />

domain, recovered by imposing Dirichlet (respectively Neumann) boundary cond<strong>it</strong>ions,<br />

provides the Neumann (respectively Dirichlet) data to be used in the contiguous do-<br />

mains. Although only half of the blocks can be simultaneously processed at any step, a<br />

speed up of the convergence is observed when using an appropriate relaxation scheme.<br />

The procedure has been investigated in zanolli(1987), funaro, quarteroni and<br />

zanolli(1988), and interpreted in quarteroni and sacchi-landriani(1988) as a pre-<br />

cond<strong>it</strong>ioned <strong>it</strong>erative method for the Schur complement of the matrix corresponding to<br />

(11.2.4)-(11.2.7). Other results are provided in zampieri(1989). A similar algor<strong>it</strong>hm<br />

for the discretization of first-order time-dependent differential problems is analyzed in<br />

quarteroni(1990).<br />

The interest in applying domain-decompos<strong>it</strong>ion methods, for the solution of partial<br />

differential equations, has increased in recent years, together w<strong>it</strong>h the development of<br />

parallel computers. Therefore, many other sophisticated techniques have been experi-<br />

mented. The reader will find a collection of papers and references in chan, glowinski,<br />

periaux and widlund(1989) and chan, glowinski, periaux and widlund(1990).


Domain-Decompos<strong>it</strong>ion Methods 263<br />

11.4 Overlapping multidomain methods<br />

To avoid complex s<strong>it</strong>uations and cumbersome notation, we confine ourselves to the case<br />

of two subdomains only. These are given by S1 :=]s0,s2[ and S2 :=]s1,s3[, where<br />

s0 < s1 < s2 < s3. The set S1 ∩ S2 =]s1,s2[ = ∅ is the overlapping region. Let us<br />

consider problem (11.2.1)(m = 3), which is equivalent to wr<strong>it</strong>ing<br />

(11.4.1)<br />

⎧<br />

⎨<br />

−U ′′<br />

1 = f in S1,<br />

⎩<br />

U1(s0) = σ1, U1(s2) = U2(s2),<br />

⎧<br />

⎨<br />

−U ′′<br />

2 = f in S2,<br />

⎩<br />

U2(s1) = U1(s1), U1(s3) = σ2,<br />

where U1 and U2 are the restrictions of U to the intervals ¯ S1 and ¯ S2 respectively.<br />

The functions U1 and U2 coincide in S1 ∩ S2.<br />

W<strong>it</strong>h the intent of using the collocation method, in analogy w<strong>it</strong>h (11.2.3), we first<br />

define the collocation nodes<br />

(11.4.2) θ (n,k)<br />

j<br />

:= 1<br />

<br />

(sk+1 − sk−1)η<br />

2<br />

(n)<br />

j<br />

+ sk+1 + sk−1<br />

<br />

0 ≤ j ≤ n, 1 ≤ k ≤ 2.<br />

For any n ≥ 2, we would now like to determine two polynomials pn,k ∈ Pn, 1 ≤ k ≤ 2,<br />

such that<br />

(11.4.3)<br />

(11.4.4)<br />

⎧<br />

⎨<br />

−p ′′ n,1(θ (n,1)<br />

i ) = f(θ (n,1)<br />

i ) 1 ≤ i ≤ n − 1,<br />

⎩<br />

pn,1(s0) = σ0, pn,1(s2) = pn,2(s2),<br />

⎧<br />

⎨<br />

−p ′′ n,2(θ (n,2)<br />

i ) = f(θ (n,2)<br />

i ) 1 ≤ i ≤ n − 1,<br />

⎩<br />

pn,2(s1) = pn,1(s1), pn,2(s3) = σ2.<br />

The value pn,1(s1) is obtained from the values pn,1(θ (n,1)<br />

i ), 0 ≤ i ≤ n, by interpolation<br />

(see formula (3.2.7)). The same argument holds for the value pn,2(s2). The first step<br />

is to show that limn→+∞ pn,k = Uk in ¯ Sk, 1 ≤ k ≤ 2. This result is proven in<br />

canuto and funaro(1988) in the Legendre case for any choice of the sk’s, and in<br />

the Chebyshev case when the size of the overlapping region is sufficiently large. Note


264 Polynomial Approximation of Differential Equations<br />

that, for any n ≥ 2, the polynomials pn,1 and pn,2 do not coincide in S1 ∩ S2. The<br />

construction of the linear system relative to (11.4.3)-(11.4.4), and the generalization to<br />

the case of more domains, is left to the reader.<br />

Finally, as in the previous section, we suggest an <strong>it</strong>erative algor<strong>it</strong>hm for the solution<br />

of the linear system corresponding to (11.4.3)-(11.4.4). For any y ∈ R, we construct<br />

two polynomials qn,k ∈ Pn, 1 ≤ k ≤ 2, satisfying<br />

(11.4.5)<br />

(11.4.6)<br />

⎧<br />

⎨ −q ′′<br />

n,1(θ (n,1)<br />

i ) = f(θ (n,1)<br />

i ) 1 ≤ i ≤ n − 1,<br />

⎩<br />

qn,1(s0) = σ0, qn,1(s2) = y,<br />

⎧<br />

⎨ −q ′′<br />

n,2(θ (n,2)<br />

i ) = f(θ (n,2)<br />

i ) 1 ≤ i ≤ n − 1,<br />

⎩<br />

qn,2(s1) = qn,1(s1), qn,2(s3) = σ2.<br />

Thus, we define the mapping Γ : R → R such that Γ(y) := qn,2(s2). Therefore,<br />

qn,k ≡ pn,k in ¯ Sk, 1 ≤ k ≤ 2, if and only if Γ(y) = y. The aim is now to find<br />

the fixed point y ∗ of the function Γ. This can be determined by the recursion formula<br />

ym+1 = Γ(ym), m ∈ N, where y0 is given. At any step, we have to solve the collocation<br />

problems (11.4.5) and (11.4.6). Such a procedure is known as the Schwarz alternating<br />

method (see schwarz(1890), Vol.2). In canuto and funaro(1988), the authors prove<br />

that limn→+∞ ym = y ∗ , for Legendre and Chebyshev approximations. In particular, in<br />

the Legendre case, we get the relation<br />

(11.4.7) |ym − y ∗ | ≤ κm<br />

1 − κ |y1 − y0| m ∈ N, where κ := (s1 − s0)(s3 − s2)<br />

(s2 − s0)(s3 − s1) .<br />

Note that the parameter κ < 1 approaches 1 when the size of the overlapping region<br />

tends to zero.<br />

A presentation of the Schwarz alternating method in a very general framework is<br />

given in dryja and widlund(1990), where a unifying theory provides a link between<br />

overlapping and non-overlapping multidomain methods.


12<br />

EXAMPLES<br />

In this chapter, we show how to apply spectral methods to the approximation of four<br />

upgraded model problems, w<strong>it</strong>hout pretending that our approach is compet<strong>it</strong>ive or im-<br />

proves upon other known numerical algor<strong>it</strong>hms. No theoretical convergence analysis is<br />

provided for the schemes proposed here.<br />

12.1 A free boundary problem<br />

In our first example, given T > 0, we search for two functions U : [−1,1] ×[0,T] → R,<br />

and Γ : [0,T] → [0,1], coupled by the following set of equations:<br />

(12.1.1)<br />

(12.1.2)<br />

∂U<br />

∂t (x,t) = ∂2U (x,t), ∀x ∈] − 1,Γ(t)[, ∀t ∈]0,T],<br />

∂x2 ∂U<br />

(−1,t) = 0, U(Γ(t),t) = 0, ∀t ∈]0,T],<br />

∂x<br />

(12.1.3) U(x,t) = 0, ∀x ∈]Γ(t),1], ∀t ∈]0,T],<br />

(12.1.4) U(x,0) = −1, ∀x ∈ [−1,0[, U(x,0) = 0, ∀x ∈ [0,1],


266 Polynomial Approximation of Differential Equations<br />

(12.1.5) Γ(0) = 0, λ Γ ′ (t) = ∂U<br />

(Γ(t),t), ∀t ∈]0,T].<br />

∂x<br />

In (12.1.5), λ is a pos<strong>it</strong>ive given constant.<br />

Let us provide a physical interpretation. We assume that the segment [−1,1]<br />

represents a one-dimensional bar of some prescribed material. In<strong>it</strong>ially, the interval<br />

[−1,0[ is in the solid state, while the interval [0,1] is in the liquid state. After appropriate<br />

scaling, the function U represents the temperature of the bar, in such a way that U = 0<br />

corresponds to the solidification temperature of the liquid. Below this level the material<br />

changes <strong>it</strong>s physical state. The in<strong>it</strong>ial configuration is expressed by cond<strong>it</strong>ions (12.1.4).<br />

Then, as the time passes, due to the lower temperature of the solid part, the liquid begins<br />

a solidification process and the boundary between the solid and the liquid regions is<br />

determined by the function Γ. For any t ∈]0,T], in the interval ] − 1,Γ(t)[ the<br />

temperature is governed by the heat equation (12.1.1) (see also section 10.2). In the<br />

remaining portion of the bar, the temperature is assumed to be zero (cond<strong>it</strong>ion (12.1.3)).<br />

Relations (12.1.2) are the boundary cond<strong>it</strong>ions of equation (12.1.1). In particular, at the<br />

endpoint x = −1 the heat flux is set to zero. Thus, the bar is thermally insulated, i.e.,<br />

there is no heat transfer across that point. Finally, the differential relation in (12.1.5)<br />

describes the evolution of Γ. This is known as the Stefan cond<strong>it</strong>ion and states that, for<br />

any t ∈]0,T], the moving point Γ(t) has a veloc<strong>it</strong>y proportional to the heat flux at<br />

(Γ(t),t). The parameter λ is the Stefan constant and <strong>it</strong> is proportional to the latent<br />

heat of the material. For high values of λ solidification is slow, and vice versa. For<br />

simplic<strong>it</strong>y, we do not take into account the effects of heat conduction inside the portion<br />

of liquid material.<br />

According to the l<strong>it</strong>erature, (12.1.1)-(12.1.5) is a single-phase Stefan problem and rep-<br />

resents an example of free boundary problem (see for instance friedman(1959), rubin-<br />

stein(1971), fasano and primicerio(1972), elliott and ockendon(1982), hill and<br />

dewynne(1987), chapter 7). A theoretical analysis shows that U and Γ are uniquely<br />

determined and Γ, as expected, is an increasing function.


Examples 267<br />

In add<strong>it</strong>ion, one can deduce that<br />

(12.1.6)<br />

= λΓ ′ (t) −<br />

= ∂U<br />

(Γ(t),t) −<br />

∂x<br />

<br />

d<br />

λ Γ(t) −<br />

dt<br />

Γ(t)<br />

−1<br />

Γ(t)<br />

−1<br />

Γ(t)<br />

This shows that the quant<strong>it</strong>y λΓ(t) − Γ(t)<br />

−1<br />

balance cond<strong>it</strong>ion).<br />

−1<br />

U(x,t) dx<br />

<br />

∂U<br />

∂t (x,t) dx − Γ′ (t) U(Γ(t),t)<br />

∂2U ∂U<br />

(x,t) dx = (−1,t) = 0, ∀t ∈]0,T].<br />

∂x2 ∂x<br />

U(x,t)dx is constant for t ∈]0,T] (heat<br />

We now propose an approximation scheme. We use spectral methods for the space<br />

variable and fin<strong>it</strong>e-differences for the time variable. We divide the interval [0,T] in<br />

m ≥ 1 parts of size h := 1/m. Then, the function Γ is approximated at the points<br />

tk := kh, 0 ≤ k ≤ m, by some values γk ∈ R, 0 ≤ k ≤ m, to be determined later.<br />

We assume that the γk’s are increasing and γ0 = 0. At the time tk, 0 ≤ k ≤ m, the<br />

function U is approximated in the interval [−1,γk] by a polynomial p (k)<br />

n ∈ Pn. On<br />

the other hand, according to (12.1.3), the approximating function is equal to zero in the<br />

interval ]γk,1]. The polynomial p (k)<br />

n , 0 ≤ k ≤ m, is defined by <strong>it</strong>s value at the n + 1<br />

points<br />

(12.1.7) θ (n,k)<br />

j<br />

Here, η (n)<br />

0<br />

:= 1<br />

(n)<br />

2 η j + 1 (1 + γk) − 1, 0 ≤ j ≤ n, 0 ≤ k ≤ m.<br />

= −1, η(n) n = 1, and η (n)<br />

j , 1 ≤ j ≤ n − 1, are the zeroes of the derivative<br />

of the Legendre polynomial of degree n (see section 3.1). In practice, the Legendre<br />

Gauss-Lobatto nodes are mapped to the interval [−1,γk] (see (11.2.3)).<br />

From (3.5.1) w<strong>it</strong>h w ≡ 1, we get the quadrature formula<br />

(12.1.8)<br />

γk<br />

−1<br />

p dx =<br />

n<br />

j=0<br />

p(θ (n,k)<br />

j ) ˜w (n,k)<br />

j<br />

∀p ∈ P2n−1, 0 ≤ k ≤ m,


268 Polynomial Approximation of Differential Equations<br />

where the new weights are defined by ˜w (n,k)<br />

j<br />

(see (3.5.6)).<br />

:= 1<br />

2 (1 + γk) ˜w (n)<br />

j , 0 ≤ j ≤ n, 0 ≤ k ≤ m<br />

We are now ready to describe the algor<strong>it</strong>hm. In view of (12.1.4), the first polynomial<br />

is determined by the relations<br />

(12.1.9) p (0)<br />

n (θ (n,0)<br />

j ) := −1, 0 ≤ j ≤ n − 1, p (0)<br />

n (θ (n,0)<br />

n ) = 0.<br />

To proceed we need to establish the new pos<strong>it</strong>ion of the free boundary. Following<br />

(12.1.5) we define<br />

(12.1.10) γk+1 := γk + h<br />

<br />

d<br />

λ dx p(k)<br />

<br />

n (γk), 0 ≤ k ≤ m − 1.<br />

The next step is to adapt the polynomial p (k)<br />

n to the stretched interval [−1,γk+1]. To<br />

this purpose we define an auxiliary polynomial q (k)<br />

n ∈ Pn satisfying for 0 ≤ j ≤ n<br />

(12.1.11)<br />

⎧<br />

⎨ q (k)<br />

⎩<br />

n (θ (n,k+1)<br />

j ) = p (k)<br />

n (θ (n,k+1)<br />

j ) if θ (n,k+1)<br />

j ∈ [−1,γk[,<br />

q (k)<br />

n (θ (n,k+1)<br />

j ) = 0 if θ (n,k+1)<br />

j ∈ [γk,γk+1].<br />

The computation of q (k)<br />

n can be done by interpolation. In fact, in analogy w<strong>it</strong>h (3.2.7),<br />

we wr<strong>it</strong>e<br />

(12.1.12) p (k)<br />

n (x) =<br />

n<br />

j=0<br />

n (θ (n,k)<br />

j ) ˜l (n,k)<br />

j (x), ∀x ∈ [−1,γk],<br />

where ˜l (n,k)<br />

j<br />

θ (n,k)<br />

i , 0 ≤ i ≤ n. It is an easy matter to check that (see also (11.2.16))<br />

(12.1.13)<br />

p (k)<br />

∈ Pn, 0 ≤ j ≤ n, are the Lagrange polynomials w<strong>it</strong>h respect to the points<br />

˜(n,k) l j (x) = ˜l (n)<br />

<br />

j 2 x+1<br />

1+γk<br />

<br />

− 1 , ∀x ∈ [−1,γk].<br />

Thus, by virtue of (3.2.8) w<strong>it</strong>h α = β = 0, we can determine the values of q (k)<br />

n at the<br />

points θ (n,k+1)<br />

j , 0 ≤ j ≤ n.


Examples 269<br />

Then, we construct the polynomial p (k+1)<br />

n<br />

consider the set of equations<br />

(12.1.14)<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

p (k+1)<br />

n<br />

(θ (n,k+1)<br />

<br />

i ) = q (k)<br />

n + h d2<br />

dx<br />

p (k+1)<br />

<br />

n (−1) = q (k)<br />

n + h d2<br />

dx<br />

by discretizing the heat equation. Thus, we<br />

p(k+1)<br />

2 n<br />

p(k+1)<br />

2 n<br />

<br />

(θ (n,k+1)<br />

i ), 1 ≤ i ≤ n,<br />

+ h[ ˜w (n,k+1)<br />

0 ] −1 d<br />

dx p(k+1)<br />

<br />

n (−1).<br />

Equation (12.1.1) has been collocated at all the points θ (n,k+1)<br />

i , 0 ≤ i ≤ n, and for<br />

i = 0 we imposed the boundary cond<strong>it</strong>ion as suggested in (9.4.23). This choice will<br />

allow us to obtain a discrete counterpart of relation (12.1.6).<br />

Basically, we used the backward Euler method (see (10.6.5)). Hence, the scheme is<br />

implic<strong>it</strong> and p (k+1)<br />

n<br />

is recovered by solving a linear system. To determine the corre-<br />

sponding matrix, we first note that, by (12.1.12) and (12.1.13), one has<br />

(12.1.15)<br />

<br />

d<br />

dx p(k+1)<br />

<br />

n<br />

=<br />

(θ (n,k+1)<br />

i ) =<br />

2<br />

1 + γk+1<br />

n<br />

j=1<br />

˜d (1)<br />

n<br />

j=0<br />

p (k+1)<br />

n<br />

(θ (n,k+1)<br />

<br />

d<br />

j )<br />

dx ˜l (n,k+1)<br />

<br />

j<br />

ij p(k+1) n (θ (n,k+1)<br />

j ), 0 ≤ i ≤ n,<br />

(θ (n,k+1)<br />

i )<br />

where we use the notations of section 7.2. Therefore, the matrix takes the form<br />

In −<br />

4h<br />

(1 + γk+1) 2 ˜ D 2 n −<br />

2h<br />

˜w (n,k+1)<br />

0 (1 + γk+1)<br />

⎡<br />

⎢<br />

⎣<br />

˜d (1)<br />

00<br />

0<br />

·<br />

·<br />

·<br />

·<br />

·<br />

·<br />

d ˜(1) 0n<br />

0<br />

· ·<br />

· ·<br />

0 · · · 0<br />

where In is the (n + 1) × (n + 1) ident<strong>it</strong>y matrix and ˜ D 2 n is the second derivative<br />

matrix at the collocation points (see section 7.2).<br />

Once we obtain p (k+1)<br />

n , we can finally update the value of the free boundary location<br />

in (12.1.10) w<strong>it</strong>h the help of (12.1.15) for i = n.<br />

⎤<br />

⎥<br />

⎦ ,


270 Polynomial Approximation of Differential Equations<br />

Provided h is sufficiently small, this algor<strong>it</strong>hm gives stable and consistent results.<br />

The restriction on h is severe (h ≈ 1/n 4 ) due to the jump in the in<strong>it</strong>ial cond<strong>it</strong>ion, but<br />

can be relaxed (h ≈ 1/n 2 ) when the heat flux at the free boundary starts decreasing in<br />

magn<strong>it</strong>ude.<br />

Numerical results for T = 1, λ = 5, n = 6, h = 1/800, are now presented to<br />

illustrate the algor<strong>it</strong>hm. The polynomials p (k)<br />

n , k = 2, 10, 100, 200, 400, 800, are<br />

plotted in figure 12.1.1. Some terms of the sequence {γk}0≤k≤800 are given in figure<br />

12.1.2.<br />

Figure 12.1.1 - Heat propagation for Figure 12.1.2 - Approximation of the<br />

the Stefan problem. free boundary.<br />

As predicted by asymptotic estimates we have Γ(t) ≈ σ √ t, where σ > 0 is related to<br />

λ. As the reader can see in figure 12.1.2, this behavior is maintained by the approximate<br />

free boundary.<br />

We conclude by noting that a relation similar to (12.1.6) holds for the discretizing<br />

functions. Actually, using (12.1.8), we have


Examples 271<br />

(12.1.16)<br />

=<br />

=<br />

=<br />

<br />

d<br />

dx p(k)<br />

<br />

n (γk) −<br />

<br />

d<br />

λ Γ(t) −<br />

dt<br />

Γ(t)<br />

−1<br />

γk<br />

U(x,t) dx<br />

≈ λ γk+1 − γk<br />

−<br />

h<br />

1<br />

[p<br />

h −1<br />

(k)<br />

<br />

d<br />

dx p(k)<br />

<br />

n (γk) − 1<br />

n<br />

[p<br />

h<br />

(k)<br />

<br />

d<br />

dx p(k)<br />

<br />

n (γk) −<br />

γk<br />

−1<br />

d 2<br />

n<br />

2 d<br />

j=0<br />

p(k)<br />

dx2 n<br />

j=0<br />

p(k)<br />

dx2 n<br />

<br />

<br />

(x) dx −<br />

<br />

t=tk<br />

n − q (k−1)<br />

n<br />

](x) dx<br />

n − q (k−1)<br />

n ](θ (n,k)<br />

j ) ˜w (n,k)<br />

j<br />

(θ (n,k)<br />

j ) ˜w (n,k)<br />

j<br />

−<br />

<br />

d<br />

dx p(k)<br />

<br />

n (−1)<br />

<br />

d<br />

dx p(k)<br />

<br />

n (−1) = 0, 1 ≤ k ≤ m − 1.<br />

A version of the scheme similar to the one proposed here has been tried for the two-phase<br />

Stefan problem in rønquist and patera(1987). Different discretizations w<strong>it</strong>h fin<strong>it</strong>e-<br />

differences for both the space and the time variable are compared in furzeland(1980).<br />

For a survey of numerical methods for free boundary problems we refer the reader to<br />

nochetto(1990).<br />

12.2 An example in an unbounded domain<br />

We study in this section the approximation of a second-order differential equation de-<br />

fined in I ≡]0,+∞[. Let f : I → R be a given function. We are concerned w<strong>it</strong>h<br />

finding the solution U : Ī → R to the problem<br />

⎧<br />

⎨<br />

(12.2.1)<br />

⎩U(0)<br />

= 0, lim U(x) = 0.<br />

x→+∞<br />

−(e x U ′ (x)) ′ = f(x) x ∈ I,<br />

We assume for example that f is such that the unique solution of (12.2.1) is U(x) =<br />

e−x sin 7x<br />

1+x2 , x ∈ Ī, which is shown in figure 12.2.1 for x ∈ [0,4].


272 Polynomial Approximation of Differential Equations<br />

Figure 12.2.1 - The function U(x) = e −x sin 7x<br />

1+x 2 , x ∈ [0,4].<br />

In our first experiment, we discretize the equation in a bounded subset of I, i.e.,<br />

the interval ]0,2[. For any n ≥ 2, we use the collocation method at the nodes ˆη (n)<br />

j :=<br />

η (n)<br />

j +1, 0 ≤ j ≤ n, obtained by shifting the Chebyshev Gauss-Lobatto points in [−1,1]<br />

(see (3.1.11)). Thus, we want to find a polynomial pn ∈ Pn such that<br />

(12.2.2)<br />

⎧<br />

⎨<br />

−p ′′ n(ˆη (n)<br />

i ) − p ′ n(ˆη (n)<br />

i ) = exp(−ˆη (n)<br />

i ) f(ˆη (n)<br />

i ), 1 ≤ i ≤ n − 1,<br />

⎩<br />

pn(0) = 0, pn(2) = 0.<br />

To determine a unique solution of (12.2.2), we force the homogeneous Dirichlet boundary<br />

cond<strong>it</strong>ion at the point x = 2. This cond<strong>it</strong>ion is suggested by the behavior prescribed<br />

for U at infin<strong>it</strong>y. We note that U(2) ≈ −.04533, so that the lim<strong>it</strong> limn→+∞ pn does<br />

not coincide w<strong>it</strong>h the function U. We examine for instance the approximation of the<br />

quant<strong>it</strong>y U ′ (0) = 7. We give in table 12.2.1 the error En := |U ′ (0) − p ′ n(0)|, n ≥ 2,<br />

for different n. As expected En does not decay to zero, but approaches a value of the<br />

same order of magn<strong>it</strong>ude of the error |U(2) − pn(2)|. Of course, we can improve on<br />

these results by increasing the size of the computational domain.


Examples 273<br />

n En<br />

4 2.88462<br />

8 .164540<br />

12 .087322<br />

16 .049646<br />

20 .052595<br />

Table 12.2.1 - Errors for the approximation<br />

of problem (12.2.1) by the scheme (12.2.2).<br />

Another approach is to use Laguerre functions (see section 6.7) for obtaining a<br />

global approximation of U in the whole domain I (see section 9.5). It is recommended<br />

not to use high degree Laguerre polynomials, due to the effect of rounding errors in<br />

the computations. Thus, we adopt a non-overlapping multidomain method (see section<br />

11.2). For any n ≥ 2 and m ≥ 1, we approximate the solution of (12.2.1) in S1 :=]0,2[<br />

by a polynomial pn ∈ Pn , and in S2 :=]2,+∞[ by a Laguerre function Qm ∈ Sm−1.<br />

At the interface point x = 2 we assume the continu<strong>it</strong>y of the approximating function<br />

and <strong>it</strong>s derivative. Then, we define the nodes ˜η (m)<br />

j := η (m)<br />

j + 2, 0 ≤ j ≤ m − 1, where<br />

η (m)<br />

0 := 0 and η (m)<br />

d<br />

j , 1 ≤ j ≤ m−1, are the zeroes of dxL(α) m for α = 0. The following<br />

collocation scheme is considered:<br />

⎧<br />

(12.2.3)<br />

−p ′′ n(ˆη (n)<br />

i ) − p ′ n(ˆη (n)<br />

i ) = exp(−ˆη (n)<br />

i ) f(ˆη (n)<br />

i ), 1 ≤ i ≤ n − 1,<br />

⎪⎨<br />

pn(0) = 0,<br />

⎪⎩<br />

pn(2) = Qm(2), p ′ n(2) = Q ′ m(2),<br />

−Q ′′ m(˜η (m)<br />

i ) − Q ′ m(˜η (m)<br />

i ) = exp(−˜η (m)<br />

i ) f(˜η (m)<br />

i ), 1 ≤ i ≤ m − 1.<br />

The subst<strong>it</strong>ution qm(x) := Qm(x)e x−2 , ∀x ∈ ¯ S2, leads to qm ∈ Pm−1, and (12.2.3)<br />

becomes<br />

(12.2.4)<br />

⎧<br />

−p ′′ n(ˆη (n)<br />

i ) − p ′ n(ˆη (n)<br />

i ) = exp(−ˆη (n)<br />

i ) f(ˆη (n)<br />

i ), 1 ≤ i ≤ n − 1,<br />

⎪⎨<br />

pn(0) = 0,<br />

⎪⎩<br />

pn(2) = qm(2), p ′ n(2) = q ′ m(2) − qm(2),<br />

−q ′′ m(˜η (m)<br />

i ) + q ′ m(˜η (m)<br />

i ) = e −2 f(˜η (m)<br />

i ), 1 ≤ i ≤ m − 1.


274 Polynomial Approximation of Differential Equations<br />

Now the error En,m := |U ′ (0) − p ′ n(0)|, depends also on the discretization parameter<br />

m. One can prove that limn,m→+∞ En,m = 0. We give in table 12.2.2 some computed<br />

values. As the reader will note, w<strong>it</strong>h nearly the same number of collocation nodes as<br />

the previous algor<strong>it</strong>hm, one may obtain better results.<br />

n En,3 En,4 En,5<br />

8 .087622 .115119 .126074<br />

12 .057295 .029799 .018844<br />

16 .031526 .004030 .006925<br />

Table 12.2.2 - Errors for the approximation<br />

of problem (12.2.1) by the scheme (12.2.4).<br />

12.3 The nonlinear Schrödinger equation<br />

Many phenomena in plasma physics, quantum physics and optics are described by the<br />

nonlinear Schrödinger equation, which is introduced for instance in trullinger, za-<br />

kharov and pokrovsky(1986). After dimensional scaling the equation takes the form<br />

(12.3.1) i ∂U<br />

∂t (x,t) + ζ ∂2 U<br />

∂x 2 (x,t) + ǫ |U(x,t)|2 U(x,t) = 0 x ∈ I, t ∈]0,T],<br />

where i = √ −1 is the complex un<strong>it</strong>y, ζ > 0 is proportional to the so called dispersion<br />

parameter, ǫ ∈ R and T > 0. The unknown U : I × [0,T] → C is a function<br />

assuming complex values. Generally we have I ≡ R and U is required to decay to zero<br />

at infin<strong>it</strong>y. For simplic<strong>it</strong>y, we take I =] − 1,1[ and impose the homogeneous boundary


Examples 275<br />

cond<strong>it</strong>ions U(±1,t) = 0, t ∈ [0,T]. Finally, an in<strong>it</strong>ial cond<strong>it</strong>ion U(x,t) = U0(x), x ∈ Ī,<br />

is provided. A conservation property is immediately obtained by noting that<br />

(12.3.2)<br />

<br />

d<br />

|U|<br />

dt I<br />

2 (x,t) dx =<br />

<br />

= i ζ U<br />

I<br />

∂2Ū ∂x2 + ǫ |U|4 − ζ Ū ∂2U ∂x<br />

<br />

I<br />

<br />

U ∂Ū<br />

∂t<br />

2 − ǫ |U|4<br />

<br />

∂U<br />

+ Ū (x,t) dx<br />

∂t<br />

<br />

(x,t) dx = 0, ∀t ∈]0,T],<br />

where the upper bar denotes the complex conjugate and we used integration by parts.<br />

This means that the quant<strong>it</strong>y <br />

I |U|2 (x,t)dx is constant w<strong>it</strong>h respect to t. Another<br />

quant<strong>it</strong>y which does not vary w<strong>it</strong>h respect to t is <br />

ζ <br />

∂U <br />

∂x<br />

2 − ǫ<br />

2 |U|4<br />

<br />

(x,t)dx.<br />

We denote by V and W the real and imaginary parts of U respectively, hence<br />

U = V + iW. Thus, we rewr<strong>it</strong>e (12.3.1) as a system<br />

(12.3.3)<br />

w<strong>it</strong>h<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

− ∂W<br />

∂t (x,t) = −ζ ∂2 V<br />

∂x 2 (x,t) − ǫ |U(x,t)|2 V (x,t)<br />

∂V<br />

∂t (x,t) = −ζ ∂2 W<br />

∂x 2 (x,t) − ǫ |U(x,t)|2 W(x,t)<br />

(12.3.4) V (±1,t) = W(±1,t) = 0, t ∈ [0,T].<br />

I<br />

x ∈ I, t ∈]0,T],<br />

The in<strong>it</strong>ial cond<strong>it</strong>ions are V (x,0) = V0(x), W(x,0) = W0(x), x ∈ Ī, U0 = V0 + iW0.<br />

For the approximation we use the collocation method based on the Legendre Gauss-<br />

Lobatto points for the variable x, and an implic<strong>it</strong> fin<strong>it</strong>e-difference scheme for the variable<br />

t (see section 10.6). The in<strong>it</strong>ial polynomials in P 0 n, n ≥ 2 (see (6.4.11)) are such that<br />

p (0)<br />

n := Ĩw,nV0 and q (0)<br />

n := Ĩw,nW0 , where the interpolation operator is defined in<br />

section 3.3 (w ≡ 1). We subdivide the interval [0,T] in m ≥ 1 equal parts of size<br />

h := T/m. The successive polynomials p (k)<br />

n , q (k)<br />

n ∈ P 0 n, 1 ≤ k ≤ m, are determined at<br />

the nodes η (n)<br />

i , 1 ≤ i ≤ n − 1, by the formulas


276 Polynomial Approximation of Differential Equations<br />

(12.3.5)<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

− q(k) n − q (k−1)<br />

n<br />

h<br />

p (k)<br />

n − p (k−1)<br />

n<br />

h<br />

= − ζ<br />

2 [p(k)<br />

= − ζ<br />

2 [q(k)<br />

n + p (k−1)<br />

n<br />

n + q (k−1)<br />

n<br />

] ′′ − ǫ<br />

4 [Θ(k) n + Θ (k−1)<br />

n<br />

] ′′ − ǫ<br />

4 [Θ(k) n + Θ (k−1)<br />

n<br />

where Θ (k)<br />

n := [p (k)<br />

n ] 2 + [q (k)<br />

n ] 2 , 0 ≤ k ≤ m. The polynomials p (m)<br />

n<br />

][p (k)<br />

][q (k)<br />

n + p (k−1)<br />

n<br />

n + q (k−1)<br />

n<br />

]<br />

and q (m)<br />

n<br />

expected to be the approximations of V (·,T) and W(·,T) respectively. Using the<br />

notations of section 9.9, (12.3.5) is equivalent to<br />

(12.3.6)<br />

⎡<br />

⎣ Mn<br />

−µIn<br />

+ ǫ<br />

⎡<br />

⎣<br />

2ζ<br />

∆(k) n<br />

where ∆ (k)<br />

n := diag (Θ (k)<br />

⎤⎡<br />

µIn<br />

⎦<br />

Mn<br />

0<br />

0 ∆ (k)<br />

n<br />

n +Θ (k−1)<br />

n<br />

⎣ ¯p(k) n<br />

¯q (k)<br />

n<br />

⎤⎡<br />

⎦<br />

⎤ ⎡<br />

⎦ = −⎣<br />

Mn<br />

⎤⎡<br />

−µIn<br />

⎦<br />

⎣ ¯p(k) n + ¯p (k−1)<br />

n<br />

¯q (k)<br />

n + ¯q (k−1)<br />

n<br />

)(η (n)<br />

1 ), · · · ,(Θ (k)<br />

µIn<br />

⎤<br />

Mn<br />

⎣ ¯p(k−1) n<br />

¯q (k−1)<br />

n<br />

⎦ 1 ≤ k ≤ m,<br />

n +Θ (k−1)<br />

n<br />

⎤<br />

⎦<br />

]<br />

are<br />

)(η (n)<br />

n−1 ) and µ := 2/ζh.<br />

The inverse of the matrix on the left-hand side of (12.3.6) can be evaluated once and<br />

for all w<strong>it</strong>h the help of (9.9.4).<br />

For any 1 ≤ k ≤ m, we can solve the nonlinear equation (12.3.6) <strong>it</strong>eratively. One<br />

approach is to construct a sequence of vectors according to the prescription<br />

(12.3.7)<br />

where ∆ (k,j)<br />

n<br />

Θ (k,j)<br />

n<br />

⎡<br />

⎣ ¯p(k,j) n<br />

¯q (k,j)<br />

n<br />

+ ǫ<br />

2ζ<br />

⎡<br />

⎤<br />

⎦ :=<br />

⎡<br />

⎣ ∆(k,j−1) n<br />

⎣ Mn<br />

:= diag (Θ (k,j)<br />

n<br />

−µIn<br />

0<br />

0 ∆ (k,j−1)<br />

n<br />

⎤−1<br />

⎧<br />

µIn ⎨<br />

⎦<br />

⎩<br />

Mn<br />

−<br />

⎡<br />

⎣ Mn<br />

⎤⎡<br />

−µIn<br />

⎦<br />

µIn Mn<br />

⎤⎡<br />

:= [p (k,j)<br />

n ] 2 + [q (k,j)<br />

n ] 2 , j ∈ N, and ¯p (k,0)<br />

n<br />

⎦<br />

⎣ ¯p(k,j−1) n<br />

¯q (k,j−1)<br />

n<br />

+ ¯p (k−1)<br />

n<br />

+ ¯q (k−1)<br />

n<br />

+ Θ (k−1)<br />

n )(η (n)<br />

1 ), · · · ,(Θ (k,j)<br />

n<br />

:= ¯p (k−1)<br />

n , ¯q (k,0)<br />

n<br />

⎤⎫<br />

⎬<br />

⎦<br />

⎭<br />

⎣ ¯p(k−1) n<br />

¯q (k−1)<br />

n<br />

j ≥ 1,<br />

⎤<br />

⎦<br />

+ Θ (k−1)<br />

n )(η (n)<br />

n−1 ) w<strong>it</strong>h<br />

:= ¯q (k−1)<br />

n .


Examples 277<br />

Therefore, one obtains<br />

(12.3.8) lim<br />

⎡<br />

⎣<br />

j→+∞<br />

¯p(k,j) n<br />

¯q (k,j)<br />

n<br />

⎤<br />

⎦ =<br />

⎡<br />

⎣ ¯p(k) n<br />

¯q (k)<br />

n<br />

⎤<br />

⎦, 1 ≤ k ≤ m.<br />

The scheme considered here was suggested in delfour, fortin and payre(1981)<br />

for fin<strong>it</strong>e-difference approximations. It is second-order accurate, i.e., the error in the<br />

variable t decays as h 2 . In add<strong>it</strong>ion, <strong>it</strong> preserves certain physical quant<strong>it</strong>ies. For<br />

example, using formula (3.5.1) for w ≡ 1, we can prove the discrete counterpart of<br />

(12.3.2), i.e., that the sum n i=0 Θ(k) n (η (n)<br />

i ) ˜w (n)<br />

i<br />

does not change w<strong>it</strong>h k.<br />

Figure 12.3.1 - Evolution of an approximate solution of equation (12.3.1).


278 Polynomial Approximation of Differential Equations<br />

We give results from a numerical experiment corresponding to the following data:<br />

ζ = 1, ǫ = 200, T = .32, n = 40, m = 64. The in<strong>it</strong>ial guess is U0(x) :=<br />

<br />

ǫ<br />

exp (iπx) cosh 2ζx <br />

<br />

, x ∈ I, U0(±1) = 0. The function Θ (k)<br />

n , plotted in figure<br />

12.3.1 for k = 0,8,16,24,32,40,48,56,64, approximates a travelling sol<strong>it</strong>on.<br />

Approximations of the nonlinear Schrödinger equation in R, by a non-overlapping<br />

multidomain method using Laguerre functions as in the previous section, are presented<br />

in de veronico(1991), and de veronico, funaro and reali(1991).<br />

12.4 Zeroes of Bessel functions<br />

Bessel functions frequently arise when solving boundary-value partial differential equa-<br />

tions by the method of separation of variables. For the sake of simplic<strong>it</strong>y, we only<br />

consider Bessel functions Jk w<strong>it</strong>h a pos<strong>it</strong>ive integer index k ≥ 1. These are deter-<br />

mined, up to a multiplicative constant, by the Sturm-Liouville problem (see section<br />

1.1):<br />

(12.4.1)<br />

⎧<br />

⎪⎨<br />

−(x J ′ k (x))′ −<br />

⎪⎩<br />

Jk(0) = 0.<br />

<br />

x − k2<br />

<br />

Jk(x) = 0, x > 0,<br />

x<br />

Many properties are known for this family of functions (they can be found, for example,<br />

in watson(1966)). In particular, we have |Jk(x)| ≤ 1/ √ 2, k ≥ 1, x > 0.<br />

An interesting issue is to evaluate the zeroes of Bessel functions. It turns out<br />

that Jk, k ≥ 1, has an infin<strong>it</strong>e sequence of pos<strong>it</strong>ive real zeroes z (k)<br />

j , j ≥ 1, that we<br />

assume to be in increasing order. We describe here one of the techniques available for<br />

the computation of the z (k)<br />

j ’s. We take for instance k = 1. For λ > 0, we define<br />

U(x) := J1( √ λx), x > 0. After subst<strong>it</strong>ution in (12.4.1), we find the differential equation<br />

(12.4.2)<br />

⎧<br />

⎨ − U<br />

⎩<br />

′′ (x) − 1<br />

x U ′ (x) + 1<br />

U(x) = λ U(x),<br />

x2 U(0) = 0.<br />

x > 0,


Examples 279<br />

By imposing the cond<strong>it</strong>ion U(1) = 0, <strong>it</strong> is possible to show that the eigenvalue problem<br />

(12.4.2) has a countable number of pos<strong>it</strong>ive real eigenvalues λm, m ≥ 1. Therefore,<br />

<strong>it</strong> is clear that z (1)<br />

j = λj, j ≥ 1. Our goal is to find an approximation to these<br />

eigenvalues. According to the results of section 8.6, for n ≥ 2, we discretize (12.4.2) by<br />

the eigenvalue problem<br />

⎧<br />

(12.4.3)<br />

⎪⎨<br />

−p ′′ n,m(˜η (n)<br />

i ) − [˜η (n)<br />

i ] −1 p ′ n,m(˜η (n)<br />

i )<br />

+ [˜η (n)<br />

i ] −2 pn,m(˜η (n)<br />

i ) = λn,m pn,m(˜η (n)<br />

i ) 1 ≤ i ≤ n − 1,<br />

⎪⎩<br />

pn,m(0) = λn,m pn,m(0), pn,m(1) = λn,m pn,m(1),<br />

0 ≤ m ≤ n,<br />

where pn,m ∈ Pn, 0 ≤ m ≤ n, and the collocation points are recovered from the<br />

Chebyshev Gauss-Lobatto nodes by setting ˜η (n)<br />

i := 1<br />

2 (η(n) i<br />

+1), 0 ≤ i ≤ n (see (3.1.11)).<br />

The (n + 1) × (n + 1) matrix relative to (12.4.3) can be wr<strong>it</strong>ten in the standard way<br />

(see chapter seven). Two of the corresponding n + 1 eigenvalues λn,m, 0 ≤ m ≤ n,<br />

are equal to 1 (say λn,0 = λn,n = 1). The others approach the exact eigenvalues<br />

λm, m ≥ 1. In table 12.4.1, we report the zeroes z (1)<br />

m and the computed eigenvalues<br />

for 1 ≤ m ≤ 5 (the first five zeroes of J1 are tabulated in luke(1969), Vol.2, p.233).<br />

In figure 12.4.1, we plot J1 in the window [0,18] × [−1,1]. An approximation of the<br />

zeroes of Jk, k ≥ 2, is obtained w<strong>it</strong>h the same arguments.<br />

m z (1)<br />

m<br />

λ8,m<br />

λ12,m<br />

λ16,m<br />

1 3.831705970207 3.831705650608 3.831705970200 3.831705970207<br />

2 7.015586669815 7.015773798574 7.015586712768 7.015586669820<br />

3 10.17346813506 10.16416616370 10.17347111164 10.17346813668<br />

4 13.32369193631 13.50667201292 13.32368681006 13.32369172635<br />

5 16.47063005087 15.87375296304 16.46311800444 16.47062464105<br />

Table 12.4.1 - Approximation of the zeroes of J1.


280 Polynomial Approximation of Differential Equations<br />

Figure 12.4.1 - Plot of J1(x), x ∈ [0,18].<br />

The proof of the orthogonal<strong>it</strong>y relation (which somehow justifies the interest for the<br />

Bessel zeroes):<br />

(12.4.4)<br />

1<br />

is left to the reader.<br />

0<br />

x Jk(z (k)<br />

i<br />

(k)<br />

x) Jk(z j x) dx = 0, if i = j, k ≥ 1,


13<br />

AN EXAMPLE<br />

IN TWO DIMENSIONS<br />

To illustrate the construction of polynomial approximations of the solution of a partial<br />

differential equation, we examine a simple boundary-value problem in two dimensions.<br />

13.1 Poisson’s equation<br />

Let Ω be the open square ] − 1,1[×] − 1,1[ in R 2 . We denote by ∂Ω the boundary<br />

of Ω. Given the function f : Ω → R, we want to find the solution U : ¯ Ω → R of the<br />

boundary-value problem<br />

⎧ 2 ∂ U<br />

⎪⎨ −∆U := −<br />

∂x<br />

(13.1.1)<br />

⎪⎩<br />

2 + ∂2U ∂y2 <br />

U ≡ 0 on ∂Ω,<br />

= f in Ω.<br />

The partial differential equation in (13.1.1) is known as Poisson’s equation. The symbol<br />

∆ is called the Laplace operator. There are many applications of equation (13.1.1) in<br />

the field of physics. For example, if Ω is a plane region of dielectric material surrounded<br />

by a grounded conductor, and f is the charge dens<strong>it</strong>y in Ω, then the solution U of<br />

(13.1.1) is proportional to the electrostatic potential on Ω.


282 Polynomial Approximation of Differential Equations<br />

Numerous theoretical results pertaining to (13.1.1) are available. For a preliminary<br />

introduction, we mention for instance courant and hilbert(1953), sneddon(1957),<br />

sobolev(1964), weinberger(1965). Here, we only state the following theorem.<br />

Theorem 13.1.1 - Let f be a continuous function in Ω satisfying <br />

then there exists a unique solution U ∈ C 0 ( ¯ Ω) ∩ C 1 (Ω) of (13.1.1).<br />

Ω f2 dxdy < +∞,<br />

The aim of the following sections is to suggest a technique for approximating the<br />

solution of Poisson’s equation by means of algebraic polynomials in two variables.<br />

13.2 Approximation by the collocation method<br />

In the following, for any n ∈ N, P ⋆ n denotes the space of polynomials in two variables<br />

which degree is less or equal to n for each variable. Thus, the dimension of the linear<br />

space P ⋆ n is (n + 1) 2 . For n ≥ 2, we also define<br />

(13.2.1) P ⋆,0<br />

n :=<br />

The dimension of P ⋆,0<br />

n is (n − 1) 2 .<br />

<br />

p ∈ P ⋆ n<br />

<br />

p ≡ 0 on ∂Ω .<br />

We analyze the approximation of problem (13.1.1) by the collocation method. For<br />

any n ≥ 2, let η (n)<br />

j , 0 ≤ j ≤ n, be the nodes in [−1,1] associated to the Ja-<br />

cobi Gauss-Lobatto formula (3.5.1). Then, in ¯ Ω we consider the set of points ℵn :=<br />

0≤i≤n<br />

0≤j≤n<br />

(η (n)<br />

i ,η (n)<br />

j ) . These nodes are displayed in figure 13.2.1 for n = 8 in the case<br />

α = β = − 1<br />

2 . We note that, if a polynomial in P⋆ n vanishes on ℵn ∩ ∂Ω , then <strong>it</strong><br />

belongs to P⋆,0 n . Moreover, any polynomial in P⋆,0 n<br />

values attained at the points in ℵn ∩ Ω.<br />

is uniquely determined by the


An Example in Two Dimensions 283<br />

Figure 13.2.1 - The set ℵ8 ⊂ ¯ Ω for α = β = − 1<br />

2 .<br />

Thus, we are concerned w<strong>it</strong>h finding a polynomial pn ∈ P ⋆,0<br />

n such that<br />

(13.2.2) −∆pn(η (n)<br />

i ,η (n)<br />

j ) = f(η (n)<br />

i ,η (n)<br />

j ), 1 ≤ i ≤ n − 1, 1 ≤ j ≤ n − 1.<br />

The above set of equations is equivalent to a (n − 1) 2 × (n − 1) 2 linear system. We<br />

examine for simplic<strong>it</strong>y the case n = 4, although the generalization to other values of n<br />

is straightforward. Let us define the vectors<br />

<br />

¯pn ≡ pn(η (n)<br />

1 ,η (n)<br />

1 ),pn(η (n)<br />

2 ,η (n)<br />

1 ),pn(η (n)<br />

3 ,η (n)<br />

1 ),pn(η (n)<br />

1 ,η (n)<br />

2 ),pn(η (n)<br />

2 ,η (n)<br />

2 ),<br />

¯fn ≡<br />

pn(η (n)<br />

3 ,η (n)<br />

2 ),pn(η (n)<br />

1 ,η (n)<br />

3 ),pn(η (n)<br />

2 ,η (n)<br />

3 ),pn(η (n)<br />

3 ,η (n)<br />

3 )<br />

<br />

f(η (n)<br />

1 ,η (n)<br />

1 ),f(η (n)<br />

2 ,η (n)<br />

1 ),f(η (n)<br />

3 ,η (n)<br />

1 ),f(η (n)<br />

1 ,η (n)<br />

2 ),f(η (n)<br />

2 ,η (n)<br />

2 ),<br />

f(η (n)<br />

3 ,η (n)<br />

2 ),f(η (n)<br />

1 ,η (n)<br />

3 ),f(η (n)<br />

2 ,η (n)<br />

3 ),f(η (n)<br />

3 ,η (n)<br />

3 )<br />

<br />

.<br />

<br />

,


284 Polynomial Approximation of Differential Equations<br />

Then, we can wr<strong>it</strong>e<br />

(13.2.3) −D ⋆ n ¯pn = ¯ fn,<br />

where, according to the notations of section 7.2, the matrix D ⋆ n for n = 4 turns out<br />

to be<br />

⎡<br />

⎢<br />

⎣<br />

2 ˜ d (2)<br />

11<br />

˜d (2)<br />

21<br />

˜d (2)<br />

31<br />

˜d (2)<br />

12<br />

˜d (2)<br />

11 + ˜ d (2)<br />

22<br />

˜d (2)<br />

32<br />

˜d (2)<br />

13<br />

˜d (2)<br />

12 0 0 ˜ d (2)<br />

13 0 0<br />

˜d (2)<br />

23 0 d ˜(2) 12 0 0 d ˜(2) 13<br />

˜d (2)<br />

11 + ˜ d (2)<br />

33 0 0 d ˜(2) 12 0 0 d ˜(2) 13<br />

˜d (2)<br />

21 0 0 d ˜(2) 11 + ˜ d (2)<br />

22<br />

0<br />

d ˜(2) 21 0 d ˜(2) 21<br />

0 0 ˜ d (2)<br />

21<br />

˜d (2)<br />

31<br />

˜d (2)<br />

12<br />

2 ˜ d (2)<br />

22<br />

˜d (2)<br />

32<br />

˜d (2)<br />

13<br />

0<br />

˜d (2)<br />

23 0 0<br />

˜d (2)<br />

23 0 d ˜(2) 23<br />

˜d (2)<br />

22 + ˜ d (2)<br />

33 0 0 ˜ d (2)<br />

23<br />

˜d (2)<br />

31 0 0 d ˜(2) 32 0 0 d ˜(2) 11 + ˜ d (2)<br />

33<br />

0<br />

d ˜(2) 31 0 0 d ˜(2) 32 0 d ˜(2) 21<br />

0 0 d ˜(2) 31 0 0 d ˜(2) 32<br />

˜d (2)<br />

31<br />

˜d (2)<br />

12<br />

˜d (2)<br />

22 + ˜ d (2)<br />

33<br />

˜d (2)<br />

32<br />

0<br />

˜d (2)<br />

13<br />

˜d (2)<br />

23<br />

2 ˜ d (2)<br />

33<br />

Such a matrix can be decomposed as D ⋆ n = An + KnAnKn, where An (n = 4) is the<br />

block diagonal matrix<br />

(13.2.4) An :=<br />

⎡<br />

Ân 0 0<br />

⎢ 0<br />

⎣<br />

Ân 0<br />

0 0 Ân<br />

⎤<br />

⎥<br />

⎦ , w<strong>it</strong>h Ân :=<br />

and Kn (n = 4) is the permutation matrix (note that Kn = K −1<br />

n ):<br />

⎡<br />

⎢<br />

⎣<br />

˜d (2)<br />

11<br />

˜d (2)<br />

21<br />

˜d (2)<br />

31<br />

˜d (2)<br />

12<br />

˜d (2)<br />

22<br />

˜d (2)<br />

32<br />

˜d (2)<br />

13<br />

˜d (2)<br />

23<br />

˜d (2)<br />

33<br />

⎤<br />

⎥<br />

⎦ ,<br />

⎤<br />

⎥<br />


An Example in Two Dimensions 285<br />

(13.2.5) Kn :=<br />

⎡<br />

1 0 0 0 0 0 0 0<br />

⎤<br />

0<br />

⎢0<br />

⎢<br />

⎢0<br />

⎢<br />

⎢0<br />

⎢<br />

⎢0<br />

⎢<br />

⎢0<br />

⎢<br />

⎢0<br />

⎣<br />

0<br />

0<br />

0<br />

1<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

1<br />

0<br />

1<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

1<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

1<br />

0<br />

1<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

1<br />

0<br />

0<br />

0 ⎥<br />

0 ⎥<br />

0 ⎥<br />

0 ⎥.<br />

⎥<br />

0 ⎥<br />

0 ⎥<br />

⎦<br />

0<br />

0 0 0 0 0 0 0 0 1<br />

Actually, for any 1 ≤ i ≤ n −1 and 1 ≤ j ≤ n −1, the transformation associated w<strong>it</strong>h<br />

Kn maps the point (η (n)<br />

i ,η (n)<br />

j ) into the point (η (n)<br />

j ,η (n)<br />

i ). Therefore, the matrix-vector<br />

multiplication ¯pn → An¯pn is equivalent to performing n − 1 second-order partial<br />

derivatives w<strong>it</strong>h respect to the variable x, and the multiplication ¯pn → KnAnKn¯pn<br />

corresponds to n − 1 second-order partial derivatives w<strong>it</strong>h respect to the variable y.<br />

The sum of the two contributions gives a discretization of the Laplace operator.<br />

We observe that, in the construction of the matrix D ⋆ n, we implic<strong>it</strong>ly assumed that the<br />

unknown polynomial pn vanishes on ∂Ω. This cond<strong>it</strong>ion is obtained by eliminating<br />

all the entries corresponding to the boundary nodes. The same trick was applied in<br />

(7.4.11) for n = 3 in the one-dimensional case.<br />

A theoretical analysis of the convergence of pn to U as n → +∞ can be developed.<br />

We outline the proof in the Legendre case (α = β = 0). We start by recasting (13.1.1) in<br />

a weak form (see section 9.3), which is obtained by multiplying the differential equation<br />

by a test function φ such that φ ≡ 0 on ∂Ω, and integrating on Ω. W<strong>it</strong>h the help of<br />

Green’s formula one gets<br />

(13.2.6)<br />

<br />

Ω<br />

<br />

∂U ∂φ<br />

∂x ∂x<br />

+ ∂U<br />

∂y<br />

<br />

∂φ<br />

dxdy =<br />

∂y<br />

<br />

Ω<br />

fφ dxdy =: F(φ).<br />

Similarly, problem (13.2.2) can be formulated in a variational setting. In fact, by formula<br />

(3.5.1) w<strong>it</strong>h w ≡ 1, integration by parts yields


286 Polynomial Approximation of Differential Equations<br />

(13.2.7)<br />

=<br />

= −<br />

n<br />

j=0<br />

1<br />

n<br />

j=0<br />

−1<br />

1<br />

∂pn<br />

∂x<br />

−1<br />

=<br />

n<br />

i=0<br />

j=0<br />

∂pn<br />

∂x<br />

∂φ<br />

∂x<br />

+ ∂pn<br />

∂y<br />

<br />

∂φ<br />

(x,η<br />

∂x<br />

(n)<br />

<br />

j ) dx ˜w (n)<br />

j<br />

2 ∂ pn<br />

φ (x,η<br />

∂x2 (n)<br />

<br />

j ) dx ˜w (n)<br />

j<br />

n<br />

i=0<br />

j=0<br />

= −<br />

n<br />

i=0<br />

j=0<br />

<br />

∂φ<br />

(η<br />

∂y<br />

(n)<br />

i ,η (n)<br />

j ) ˜w (n)<br />

i ˜w (n)<br />

j<br />

+<br />

−<br />

n<br />

i=0<br />

n<br />

i=0<br />

1<br />

−1<br />

1<br />

−1<br />

∂pn<br />

∂y<br />

[∆pn φ](η (n)<br />

i ,η (n)<br />

j ) ˜w (n)<br />

i ˜w (n)<br />

j<br />

<br />

∂φ<br />

(η<br />

∂y<br />

(n)<br />

<br />

i ,y) dy ˜w (n)<br />

i<br />

2 ∂ pn<br />

φ (η<br />

∂y2 (n)<br />

<br />

i ,y) dy ˜w (n)<br />

i<br />

[fφ](η (n)<br />

i ,η (n)<br />

j ) ˜w (n)<br />

i ˜w (n)<br />

j =: Fn(φ), ∀φ ∈ P ⋆,0<br />

n .<br />

Basically, the expression above follows from (13.2.6) by replacing the integrals w<strong>it</strong>h<br />

the help of a quadrature formula based on the points belonging to ℵn. Finally, we<br />

can estimate a certain norm of the error |U − pn| w<strong>it</strong>h a su<strong>it</strong>able generalization<br />

of theorem 9.4.1. The rate of convergence is obtained by examining the error due<br />

to the use of the quadrature in place of the exact integrals. For example, the error<br />

|F(φ) − Fn(φ)|, φ ∈ P ⋆,0<br />

n , converges to zero w<strong>it</strong>h a rate depending on the smoothness<br />

of the function f (see canuto, hussaini, quarteroni and zang(1988), section 9.7).<br />

As usual, the analysis of the other Jacobi cases is more complicated. The Chebyshev<br />

case (α = β = −1 2 ) is treated for instance in bressan and quarteroni(1986b). The<br />

strategy in this class of proofs is to deal w<strong>it</strong>h the two variables x and y separately, in<br />

order to take advantage of the theory developed for the one-dimensional case.<br />

Of course, alternate spectral discretizations of Poisson’s equation, such as the tau<br />

method, may be considered. Further generalizations result from considering approx-<br />

imating polynomials of different degrees w<strong>it</strong>h respect to the variables x and y. The<br />

same techniques apply to other partial differential equations defined in Ω. For exam-<br />

ple, spectral methods have been successfully applied to approximate the Navier-Stokes


An Example in Two Dimensions 287<br />

equations, which are briefly introduced in section 13.4. However, the theoretical study<br />

of convergence can be qu<strong>it</strong>e involved even for simple problems.<br />

Setting up an approximation scheme is not an easy task when Ω is not a square. In this<br />

s<strong>it</strong>uation, one can try to map Ω into a square and, after change of variables, discretize<br />

the new differential problem. Another approach, sometimes combined w<strong>it</strong>h the previous<br />

one, is to use a domain-decompos<strong>it</strong>ion method. The reader interested in this kind<br />

of applications can find a good list of references in the books of canuto, hussaini,<br />

quarteroni and zang(1988), boyd(1989).<br />

13.3 Hints for the implementation<br />

Besides the standard techniques used to numerically solve (13.2.3), other algor<strong>it</strong>hms,<br />

related to the specific structure of the matrix D ⋆ n, n ≥ 2, can be devised and imple-<br />

mented. First, let us introduce some notations. The tensor product of two n × n<br />

matrices W ≡ {wij} 1≤i≤n<br />

1≤j≤n<br />

matrix<br />

(13.3.1) W ⊗ Z :=<br />

and Z ≡ {zij} 1≤i≤n<br />

1≤j≤n<br />

⎡<br />

w11Z w12Z · · ·<br />

⎤<br />

w1nZ<br />

⎢ w21Z<br />

⎢ .<br />

⎢ .<br />

⎣<br />

w22Z<br />

.<br />

· · ·<br />

⎥<br />

w2nZ ⎥<br />

⎥.<br />

.<br />

⎥<br />

⎦<br />

wn1Z wn2Z · · · wnnZ<br />

If W and Z are invertible, then we have the relation<br />

(13.3.2) (W ⊗ Z) −1 = W −1 ⊗ Z −1 .<br />

is defined to be the n 2 × n 2 block<br />

Let În denote the (n − 1) × (n − 1) ident<strong>it</strong>y matrix. Thus, we obtain the following<br />

expression:<br />

(13.3.3) D ⋆ n = În ⊗ Ân + Ân ⊗ În,


288 Polynomial Approximation of Differential Equations<br />

where Ân is the (n − 1) × (n − 1) matrix { ˜ d (2)<br />

ij } 1≤i≤n−1<br />

1≤j≤n−1<br />

(see (13.2.4) for n = 4).<br />

In section 8.2 we claimed that for −1 < α < 1 and −1 < β < 1, the eigenvalues of<br />

− Ân are real pos<strong>it</strong>ive and distinct, although this cannot always be proven.<br />

Under this assumption, we can find an invertible (n − 1) × (n − 1) matrix Bn such<br />

that Ân = B −1<br />

n ΛnBn, where Λn is diagonal. At this point, one easily checks that<br />

(Bn ⊗ Bn) −1 D ⋆ n(Bn ⊗ Bn) is equal to the diagonal matrix În ⊗ Λn + Λn ⊗ În. This<br />

shows that the eigenvalues of −D ⋆ n can be recovered from those of problem (8.2.1). In<br />

particular, they are λn,m + λn,k, 1 ≤ m ≤ n − 1, 1 ≤ k ≤ n − 1. From (13.3.2), we<br />

finally obtain the following relation (see lynch, rice and thomas(1964)):<br />

(13.3.4) (D ⋆ n) −1 = (Bn ⊗ Bn)( În ⊗ Λn + Λn ⊗ În) −1 (B −1<br />

n ⊗ B −1<br />

n ).<br />

This suggests the possibil<strong>it</strong>y of performing the multiplication ¯ fn → (D ⋆ n) −1 ¯ fn very<br />

efficiently, provided we know how to decompose the block Ân.<br />

System (13.2.3) can be also solved by a precond<strong>it</strong>ioned <strong>it</strong>erative method (see section<br />

8.3). Following section 8.4, we can construct a (n−1)×(n−1) precond<strong>it</strong>ioning matrix<br />

ˆRn for − Ân, based on a fin<strong>it</strong>e-difference scheme at the collocation nodes. Then, one<br />

finds that the (n − 1) 2 × (n − 1) 2 matrix Rn := În ⊗ ˆ Rn + ˆ Rn ⊗ În is a good<br />

precond<strong>it</strong>ioner for −D ⋆ n, since the eigenvalues of R −1<br />

n D ⋆ n and ˆ R −1<br />

n Ân have the same<br />

qual<strong>it</strong>ative behavior. In add<strong>it</strong>ion, since Rn has the same structure of D ⋆ n, we can<br />

invert <strong>it</strong> by the procedure described above.<br />

Another <strong>it</strong>erative technique is the alternating-direction method, which is detailed<br />

in peaceman and rachford(1955), douglas and rachford(1956), varga(1962).<br />

Let us fix the the real parameter ω > 0 and denote by In the (n − 1) 2 × (n − 1) 2<br />

ident<strong>it</strong>y matrix. The algor<strong>it</strong>hm consists of two successive steps, i.e.,<br />

(13.3.5) (−An + ωIn) ¯q (k)<br />

n := ¯ fn + (KnAnKn + ωIn) ¯p (k)<br />

n , k ∈ N,<br />

(13.3.6) (−KnAnKn + ωIn) ¯p (k+1)<br />

n<br />

where the in<strong>it</strong>ial guess ¯p (0)<br />

n<br />

:= ¯ fn + (An + ωIn) ¯q (k)<br />

n , k ∈ N,<br />

is assigned and {¯q (k)<br />

n }k∈N is an auxiliary sequence of<br />

vectors. Then we have limk→+∞ ¯p (k)<br />

n = ¯pn, where ¯pn is the solution of (13.2.3).


An Example in Two Dimensions 289<br />

The scheme (13.3.5)-(13.3.6) is implic<strong>it</strong>. However, we can invert the block diagonal<br />

matrix −An + ωIn in (13.3.5) by computing once and for all the inverse of the block<br />

− Ân + ω În. Similarly, by wr<strong>it</strong>ing (−KnAnKn + ωIn) −1 = Kn(−An + ωIn) −1 Kn, we<br />

can deal w<strong>it</strong>h (13.3.6). The theory shows that the fastest convergence of the method is<br />

obtained when ω = λn,1 λn,n−1, where λn,1 and λn,n−1 are the smallest and the<br />

largest eigenvalues of − Ân respectively.<br />

To conclude this section, we present the results of a numerical test. In (13.1.1) we<br />

take f(x,y) := 2π, ∀(x,y) ∈ Ω, i.e., a constant charge dens<strong>it</strong>y in Ω. The corresponding<br />

potential U is plotted in figure 13.3.1. Besides, in table 13.3.1, we report for different n<br />

the value of the approximating polynomial pn at the center of Ω. Chebyshev collocation<br />

nodes have been used. As n increases, these computed quant<strong>it</strong>ies converge to a lim<strong>it</strong>,<br />

which is expected to be the potential at the point (0,0).<br />

Figure 13.3.1 - Solution of problem (13.1.1) for f ≡ 2π.


290 Polynomial Approximation of Differential Equations<br />

n pn(0,0)<br />

2 1.570796326<br />

4 1.856395658<br />

6 1.851713702<br />

8 1.851563407<br />

10 1.851562860<br />

12 1.851563065<br />

Table 13.3.1 - Approximations of U(0,0).<br />

13.4 The incompressible Navier-Stokes equations<br />

The Navier-Stokes equations describe, w<strong>it</strong>h a good degree of accuracy, the motion of a<br />

viscous incompressible fluid in a region Ω (see for instance batchelor(1967)). They<br />

find applications in aeronautics, meteorology, plasma physics, and in other sciences.<br />

In two dimensions Ω is an open set of R 2 . The unknowns, depending on the space<br />

variables (x,y) ∈ ¯ Ω and the time variable t ∈ [0,T], T > 0, are the two components<br />

of the veloc<strong>it</strong>y vector field V ≡ (V1,V2), Vi : Ω × [0,T] → R, 1 ≤ i ≤ 2, and the<br />

pressure P : Ω×]0,T] → R. Assuming that the dens<strong>it</strong>y of the fluid is constant, after<br />

dimensional scaling the equations become<br />

(13.4.1)<br />

(13.4.2)<br />

(13.4.3)<br />

∂V1<br />

∂t<br />

∂V2<br />

∂t<br />

∂V1<br />

+ V1<br />

∂x<br />

∂V2<br />

+ V1<br />

∂x<br />

∂V1<br />

∂x<br />

+ ∂V2<br />

∂y<br />

∂V1<br />

+ V2<br />

∂y − ν ∆V1 = − ∂P<br />

∂x<br />

∂V2<br />

+ V2<br />

∂y − ν ∆V2 = − ∂P<br />

∂y<br />

= 0 in Ω×]0,T],<br />

in Ω×]0,T],<br />

in Ω×]0,T],


An Example in Two Dimensions 291<br />

where ν > 0 is a given parameter associated w<strong>it</strong>h the kinematic viscos<strong>it</strong>y, and ∆ is<br />

the Laplace operator defined in (13.1.1). The nonlinear relations (13.4.1) and (13.4.2)<br />

are known as momentum equations, while (13.4.3) is the continu<strong>it</strong>y (or incompressibil-<br />

<strong>it</strong>y) equation. They express the conservation laws of the momentum and the mass<br />

respectively. We note the analogy between the momentum equations and the Burgers<br />

equation, presented in section 10.4 for the case of one space variable. The functions<br />

Vi, 1 ≤ i ≤ 2, require in<strong>it</strong>ial and boundary cond<strong>it</strong>ions. For example, by imposing<br />

Vi ≡ 0 on ∂Ω×]0,T], 1 ≤ i ≤ 2, we are specifying a no-slip cond<strong>it</strong>ion at the bound-<br />

ary. Instead, no in<strong>it</strong>ial or boundary cond<strong>it</strong>ions are needed for the pressure, which is<br />

determined up to an add<strong>it</strong>ive constant.<br />

For a theoretical study of the equations (13.4.1)-(13.4.3), the reader is inv<strong>it</strong>ed to<br />

consult the books of ladyzhenskaya(1969), temam(1985), kreiss and lorenz(1989)<br />

and the numerous references contained therein. On the subjects of numerical approxi-<br />

mation by fin<strong>it</strong>e-differences or fin<strong>it</strong>e element methods the bibliography is very rich, and<br />

a comprehensive list of references cannot be included here. We would like to remark<br />

that the numerical investigation of the equations relevant to the physics of fluids has<br />

also been a primary source of algor<strong>it</strong>hms and solution techniques in the field of spectral<br />

methods. Due to the abundance of results, we are unable to mention all the contribu-<br />

tors. A distinguished referring point is the book of canuto, hussaini, quarteroni<br />

and zang(1988), which is specifically addressed to computations in fluid dynamics. The<br />

reader will find there material for further extensions.<br />

* * * * * * * * * * * * * * *


REFERENCES<br />

adams r.(1975), Sobolev Spaces, Academic Press, New York.<br />

agoshkov v.i.(1988), Poincaré-Steklov’s operators and domain decompos<strong>it</strong>ion methods in fin<strong>it</strong>e dimensional<br />

spaces, Proceedings of the I International Symposium on Domain Decompos<strong>it</strong>ion Methods<br />

for P.D.E., SIAM, Philadelphia.<br />

askey r.(1975), Orthogonal Polynomials and Special Functions, SIAM, Philadelphia.<br />

aubin j.p.(1972), Approximation of Elliptic Boundary-Value Problems, Wiley-Interscience, New York.<br />

babu˘ska i. & suri m.(1987), The p-version of the fin<strong>it</strong>e element method w<strong>it</strong>h quasiuniform meshes, M 2 AN,<br />

21, pp.199-238.<br />

babu˘ska i., szabo b.a. & katz i.n.(1981), The p-version of the fin<strong>it</strong>e element method, SIAM J. Numer.<br />

Anal., 18, n.3, pp.515-545.<br />

baldock g.r.& bridgeman t.(1981), The mathematical Theory of Wave Motion, Ellis Horwood Lim<strong>it</strong>ed,<br />

Chichester (England).<br />

batchelor g.k.(1967), An introduction to Fluid Dynamics, Cambridge Univers<strong>it</strong>y Press, Cambridge.<br />

bellman r.(1944), A note on an inequal<strong>it</strong>y of E. Schmidt, Bulletin of the Amer. Math. Soc., 50, n.10,<br />

pp.734-736.<br />

bergh j. & löfström(1976), Interpolation Spaces, Spinger-Verlag, Berlin.<br />

bernardi c., deb<strong>it</strong> n. & maday y.(1990), Coupling fin<strong>it</strong>e element and spectral methods: first results,<br />

Math. Comp., 54, n.189, pp.21-39.<br />

bernardi c. & maday y.(1988), A collocation method over staggered grids for the Stokes problem,<br />

Internat. J. for Numer. Methods in Fluids, 8, pp.537-557.<br />

bernardi c. & maday y.(1989), Properties of some weighted Sobolev spaces and application to spectral<br />

approximations, SIAM J. Numer. Anal., 26, n.4, pp.769-829.<br />

bernardi c. & maday y.(1990), Some spectral approximations of monodimensional fourth-order problems,<br />

Manuscript, to appear in J. Approx. Theory.<br />

bernardi c., maday y. & sacchi-landriani g.(1989), Nonconforming matching cond<strong>it</strong>ions for coupling<br />

spectral and fin<strong>it</strong>e element methods, Applied Numer. Math., 6, pp.65-84.<br />

bernstein s.(1912a), Sur l’ordre de la meilleure approximation des fonctions continues par des polynômes<br />

de degré donné, Mém. Acad. Roy. Belgique, 2me série, 4, pp.1-104.<br />

bernstein s.(1912b), Démonstration du théorème de Weierstrass fondée sur le calcul de probabil<strong>it</strong>és, CKMS,<br />

13, pp.1-2.


294 Polynomial Approximation of Differential Equations<br />

bertoluzza s. & funaro d.(1991), Remarks about the matrices of the pseudospectral approximation of<br />

Neumann problems, in the Proceedings of the IMACS Int. Sym. on Iterative Methods in Linear<br />

Algebra, Brussels, to appear.<br />

bini d., capovani m., lotti g. & romani f.(1981), Compless<strong>it</strong>à Numerica, Boringhieri, Torino.<br />

bjørstad p.e. & widlund o.b.(1986), Iterative methods for the solution of elliptic problems on regions<br />

part<strong>it</strong>ioned into substructures, SIAM J. Numer. Anal., 23, n.6, pp.1097-1120.<br />

boyd j.p.(1980), The rate of convergence of Herm<strong>it</strong>e functions series, Math. Comp., 35, n.152, pp.1309-1316.<br />

boyd j.p.(1981), The rate of convergence of Chebyshev polynomials for functions which have asymptotic<br />

power series about one endpoint, Math. Comp., 37, n.155, pp.189-195.<br />

boyd j.p.(1982), The optimization of convergence for Chebyshev polynomial methods in an unbounded<br />

domain, J. Comput. Phys., 45, pp.43-79.<br />

boyd j.p.(1984), Asymptotic coefficients of Herm<strong>it</strong>e functions series, J. Comput. Phys., 54, n.3, pp.382-410.<br />

boyd j.p.(1989), Chebyshev & Fourier Spectral Methods, Lecture Notes in Engineering, Springer, New<br />

York.<br />

brachet m.e., meiron d.i., orszag s.a., nickel b.g., morf r.h. & frisch u.(1983), Small scale<br />

structures of the Taylor-Green vortex, J. Fluid Mech., 130, pp.411-452.<br />

brandt a., fulton s.r. & taylor g.d.(1985), Improved spectral multigrid methods for periodic elliptic<br />

problems, J. Comput. Phys., 58, pp.96-112.<br />

brauer f. & nohel j.a.(1967), Ordinary Differential Equations, W.A. Benjamin, New York.<br />

bressan n.& pavoni d.(1990), Truncation versus mapping in the spectral approximation to the Kortewegde<br />

Vries equation, in Spectral and High Order Methods for P.D.E. (C.Canuto & A.Quarteroni<br />

Eds.), North Holland, Amsterdam, pp.443-450.<br />

bressan n.& quarteroni a.(1986a), An implic<strong>it</strong>/explic<strong>it</strong> spectral method for Burgers equation, Calcolo,<br />

23, pp.265-284.<br />

bressan n.& quarteroni a.(1986b), Analysis of Chebyshev collocation methods for parabolic equations,<br />

SIAM J. Numer. Anal., 23, n.6, pp.1138-1154.<br />

brezis h.(1983), Analyse Fonctionnelle, Masson, Paris.<br />

brezzi f. & gilardi g.(1987), in Fin<strong>it</strong>e Element Handbook (H.Kardenstuncer Ed.), McGraw-Hill, New<br />

York, p.1-121.<br />

brezzi f., rappaz j. & raviart p.a.(1980), Fin<strong>it</strong>e dimensional approximation of nonlinear problems.<br />

Part I: branches of nonsingular solutions, Numer. Math., 36, pp.1-25.<br />

brigham e.o.(1974), The Fast Fourier Transform, Prentice-Hall, Englewood Cliffs NJ.<br />

cai w., gottlieb d. & harten a.(1990), Cell averaging Chebyshev methods for hyperbolic problems,<br />

ICASE (Hampton VA) report n.90-27.<br />

calogero f.(1983), Computation of Sturm-Liouville eigenvalues via Lagrangian interpolation, Lett. Nuovo<br />

Cimento, 37, pp.9-16.<br />

calogero f. & franco e.(1985), Numerical tests of a novel technique to compute the eigenvalues of<br />

differential operators, Nuovo Cimento, 89b, p.161.<br />

canuto c.(1986), Boundary cond<strong>it</strong>ions in Chebyshev and Legendre methods, SIAM J. Numer. Anal., 23,<br />

n.4, pp.815-831.<br />

canuto c.(1988), Spectral methods and a maximum principle, Math. Comp., 51, n.184, pp.615-629.<br />

canuto c. & funaro d.(1988), The Schwarz algor<strong>it</strong>hm for spectral methods, SIAM J. Numer. Anal., 25,<br />

n.1, pp.24-40.


References 295<br />

canuto c., hussaini m.y., quarteroni a.& zang t.a.(1988), Spectral Methods in Fluid Dynamics,<br />

Springer Series in Computational Physics, Springer-Verlag, New York.<br />

canuto c. & quarteroni a.(1980), Propriétés d’approximation dans les espaces de Sobolev de systémes<br />

de polinômes orthogonaux, C.R. Acad. Sci. Paris, 290, pp.A925-A928.<br />

canuto c. & quarteroni a.(1981), Spectral and pseudo-spectral methods for parabolic problems w<strong>it</strong>h<br />

nonperiodic boundary cond<strong>it</strong>ions, Calcolo, 18, pp.197-218.<br />

canuto c. & quarteroni a.(1982a), Approximation results for orthogonal polynomials in Sobolev spaces,<br />

Math. Comp., 38, pp.67-86.<br />

canuto c. & quarteroni a.(1982b), Error estimates for spectral approximations of hyperbolic equations,<br />

SIAM J. Numer. Anal., 19, pp.629-642.<br />

canuto c. & quarteroni a.(1984), Variational methods in the theoretical analysis of spectral approximations,<br />

in Spectral Methods for Partial Differential Equations (R.G. Voigt, D. Gottlieb & M.Y.<br />

Hussaini Eds.), SIAM-CBMS, Philadelphia, pp.55-78.<br />

canuto c. & quarteroni a.(1985), Precond<strong>it</strong>ioned minimal residual methods for Chebyshev spectral<br />

calculations, J. Comput. Phys., 60, pp.315-337.<br />

canuto c. & quarteroni a.(1987), On the boundary treatment in spectral methods for hyperbolic<br />

systems, J. Comput. Phys., 71, pp.100-110.<br />

chan t.f., glowinski r., periaux j. & widlund o.b. (Ed<strong>it</strong>ors) (1989), Domain Decompos<strong>it</strong>ion<br />

Methods, SIAM, Philadelphia.<br />

chan t.f., glowinski r., periaux j. & widlund o.b. (Ed<strong>it</strong>ors) (1990), Domain Decompos<strong>it</strong>ion<br />

Methods for Partial Differential Equations, SIAM, Philadelphia.<br />

chan t.f. & goovaerts d.(1989), Schur complement domain decompos<strong>it</strong>ion algor<strong>it</strong>hms for spectral<br />

methods, Applied Numer. Math., 6, pp.53-64.<br />

chang k.w. & howes f.a.(1984), Nonlinear Singular Perturbation Phenomena: Theory and Applications,<br />

Springer, New York.<br />

chatelin f.(1983), Spectral Approximation of Linear Operators, Academic Press, New York.<br />

cheney e.w.(1966), Introduction to Approximation Theory, Mc Graw-Hill, New York.<br />

ciarlet p.g.(1978), The Fin<strong>it</strong>e Element Method for Elliptic Problems, North Holland, Amsterdam.<br />

clenshaw c.w.(1957), The numerical solution of linear differential equations in Chebyshev series, Proc.<br />

Cambridge Philos. Soc., 53, pp.134-149.<br />

cooley j.w., lewis p.a. & welch p.d.(1970), The fast Fourier transform algor<strong>it</strong>hm: programming<br />

considerations in the calculations of sine, cosine and Laplace transform, J. Sound Vibration, 12,<br />

pp.315-317.<br />

cooley j.w. & tukey j.w.(1965), An algor<strong>it</strong>hm for the machine calculation of complex Fourier series,<br />

Numer. Math., 2, pp.197-205.<br />

comincioli v.(1990), Analisi Numerica, Metodi Modelli Applicazioni, McGraw-Hill Libri Italia, Milan.<br />

courant r. & hilbert d.(1953), Methods of Mathematical Physics, Wiley-Interscience, New York.<br />

dahlquist g.& björck h.b.(1974), Analysis of Numerical Methods, John Wiley & Sons, New York.<br />

davis p.j. & rabinow<strong>it</strong>z p.(1984), Methods of Numerical Integration, Second ed<strong>it</strong>ion, Academic Press,<br />

New York.<br />

delfour m., fortin m. & payre g.(1981), Fin<strong>it</strong>e-difference solutions of a non-linear Schrödinger equation,<br />

J. Comput. Phys., 44, pp.277-288.<br />

delves l.m.(1968), The numerical evaluation of principal value integrals, Comput. J., 10, pp.389-391.


296 Polynomial Approximation of Differential Equations<br />

dettman j.w.(1969), Mathematical Methods in Physics and Engineering, McGraw-Hill, New York.<br />

de veronico m.c.(1991), Approssimazione dell’equazione di Schrödinger mediante polinomi algebrici, Thesis,<br />

Univers<strong>it</strong>y of Pavia, Italy.<br />

de veronico m.c., funaro d. & reali g.c.(1991), A novel numerical technique to investigate nonlinear<br />

guided waves: approximation of nonlinear Schrödinger equation by nonperidic pseudospectral<br />

methods, pre-print n.800, Ist<strong>it</strong>uto di Analisi Numerica del C.N.R., Pavia, Italy.<br />

deville m. & mund e.(1985), Chebyshev pseudospectral solution of second-order elliptic equations w<strong>it</strong>h<br />

fin<strong>it</strong>e element pre-cond<strong>it</strong>ioning, J. Comput. Phys., 60, pp.517-533.<br />

d<strong>it</strong>zian z. & totik v.(1987), Moduli of Smoothness, Springer-Verlag, New York.<br />

douglas j. Jr. & rachford h.h. Jr.(1956), On the numerical solution of heat conduction problems in<br />

two or three space variables, Trans. Amer. Math. Soc., 82, pp.421-439.<br />

dryja m. & widlund o.b.(1990), Towards a unified theory of domain decompos<strong>it</strong>ion algor<strong>it</strong>hms for elliptic<br />

problems, in Domain Decompos<strong>it</strong>ion Methods for P.D.E. (T.F. Chan, R. Glowinski, J. Periaux<br />

& O.B. Widlund Eds.), SIAM, Philadelphia.<br />

dubiner m.(1987), Asymptotic analysis of spectral methods, J. Sci. Comput., 2, pp.3-32.<br />

duff i.s., erisman a.m. & reid j.k.(1986), Direct Methods for Sparse Matrices, Clarendon Press,<br />

Oxford.<br />

duffin r.j. & schaeffer a.c.(1941), A refinement of an inequal<strong>it</strong>y of the brothers Markoff, Trans. Amer.<br />

Math. Soc., 50, pp.517-528.<br />

durand l.(1985), Lagrangian differentiation, Gauss-Jacobi integration and Sturm-Liouville eigenvalue problems,<br />

Int. Symp. on Orthogonal Polynomials and their Applications, Bar-le-Duc (France), october<br />

1984, Lecture Notes in Mathematics, Springer-Verlag.<br />

eckhaus w.(1979), Asymptotic Analysis of Singular Perturbations, North-Holland, Amsterdam.<br />

elliott d.(1981), A convergence theorem for singular integral equations, J. Austral. Math. Soc., B22,<br />

pp.539-552.<br />

elliott d.(1982), The classical collocation method for solving singular integral equations, SIAM J. Numer.<br />

Anal., 19, pp.816-832.<br />

elliott c.m. & ockendon j.r.(1982), Weak and Variational Methods for Moving Boundary Problems,<br />

P<strong>it</strong>man, London.<br />

engels h.(1980), Numerical Quadrature and Cubature, Academic Press, London.<br />

epstein b.(1962), Partial Differential Equations, An Introduction, McGraw-Hill, New York.<br />

erdogan f., gupta g.d. & cook t.s.(1973), Numerical solution of singular integral equations, in Mechanics<br />

of Fracture, Vol.1 (G.S. Sih Ed.), Noordhoff, Leyden, pp.368-425.<br />

erdös p.(1961), Problems and results on the theory of interpolation. II, Acta Math. Acad. Sci. Hungar,<br />

12, pp.235-244.<br />

escobedo m. & kavian o.(1987), Variational problems related to self-similar solutions of the heat equation,<br />

Nonlinear Anal., 11, n.10, pp.1103-1133.<br />

fasano a. & primicerio m.(1972), Il problema di Stefan con condizioni al contorno non lineari, Ann.<br />

Scuola Norm. Sup. Pisa, 26, pp.711-737.<br />

feinerman r.p. & newman d.j.(1974), Polynomial Approximation, Waverly Press, Baltimore MD.<br />

fletcher c.a.j.(1984), Computational Galerkin Methods, Springer-Verlag, New York.<br />

fox l.(1964), An introduction to Numerical Linear Algebra, Clarendon Press, Oxford.


References 297<br />

friedman a.(1959), Free boundary problems for parabolic equations. I, melting of solids, J. Math. Mech.,<br />

8, pp.499-517.<br />

funaro d.(1986), A multidomain spectral approximation of elliptic equations, Numer. Methods Partial<br />

Differential Equations, 2, pp.187-205.<br />

funaro d.(1987), A precond<strong>it</strong>ioning matrix for the Chebyshev differencing operator, SIAM J. Numer. Anal.,<br />

24, pp.1024-1031.<br />

funaro d.(1988), Domain decompos<strong>it</strong>ion methods for pseudospectral approximations. Part I: second order<br />

equations in one dimension, Numer. Math., 52, pp.329-344.<br />

funaro d.(1990a), Computational aspects of pseudospectral Laguerre approximations, Applied Numer.<br />

Math., 6, pp.447-457.<br />

funaro d.(1990b), Convergence analysis for pseudospectral multidomain approximations of linear advection<br />

equations, IMA J. of Numer. Anal., 10, pp.63-74.<br />

funaro d.(1991), Error estimates for approximations by Laguerre polynomials in Sobolev spaces, in the<br />

Proceedings of the III Int. Sym. on Orthogonal Polynomials and their Applications (C.Brezinski,<br />

L.Gori & A.Ronveaux Eds.), J.C.Baltzer, Basel, pp.263-266.<br />

funaro d. & gottlieb d.(1988), A new method of imposing boundary cond<strong>it</strong>ions in pseudospectral<br />

approximations of hyperbolic equations, Math. Comp., 51, n.184, pp.599-613.<br />

funaro d. & gottlieb d.(1989), Convergence results for pseudospectral approximations of hyperbolic<br />

systems, ICASE (Hampton VA) report n.89-59.<br />

funaro d. & heinrichs w.(1990), Some results about the pseudospectral approximation of one-dimensional<br />

fourth-order problems, Numer. Math., 58, pp.399-418.<br />

funaro d. & kavian o.(1988), Approximation of some diffusion evolution equations in unbounded domains<br />

by Herm<strong>it</strong>e functions, Preprint of the Inst<strong>it</strong>ut Élie Cartan, n.15, Nancy (France).<br />

funaro d., quarteroni a. & zanolli p.(1988), An <strong>it</strong>erative procedure w<strong>it</strong>h interface relaxation for<br />

domain decompos<strong>it</strong>ion methods, SIAM J. Numer. Anal., 25, pp.1213-1236.<br />

funaro d. & rothman e.(1989), Precond<strong>it</strong>ioning matrices for the pseudo spectral approximation of firstorder<br />

operators, in Fin<strong>it</strong>e Element Analysis in Fluids (T.J.Chung & G.R.Karr Eds.), UAH Press,<br />

Huntsville AL, pp.1458-1463.<br />

furzeland r.m.(1980), A comparative study of numerical methods for moving boundary problems, J. Inst.<br />

Math. Appl., 26, pp.411-429.<br />

garnir h.g., de wilde m. & schmets j.(1973), Analyse Fonctionelle, Birkhäuser Verlag, Basel.<br />

gastaldi f. & quarteroni a.(1989), On the coupling of hyperbolic and parabolic systems: analytical<br />

and numerical approach, Applied Numer. Math., 6, pp.3-31.<br />

gatteschi l.(1964), Proprietà asintotiche di una funzione associata ai polinomi di Laguerre e loro utilizzazione<br />

al calcolo numerico degli zeri dei polinomi stessi, Atti Accad. Sci. Torino, 98.<br />

gautschi w. & li s.(1991), The remainder term for analytic functions of Gauss-Radau and Gauss-Lobatto<br />

quadrature rules w<strong>it</strong>h multiple end points, J. Comput. Appl. Math., to appear.<br />

gear c.w.(1971), Numerical In<strong>it</strong>ial Value Problems in Ordinary Differential Equations, Prentice-Hall,<br />

Englewood Cliffs NJ.<br />

ghizzetti a. & ossicini a.(1970), Quadrature Formulae, Academic Press, New York.<br />

goffman c. & pedrick g.(1965), First Course in Functional Analysis, Prentice-Hall, Englewood Cliffs<br />

NJ.<br />

golomb m. & shanks m.(1965), Elements in Ordinary Differential Equations, McGraw-Hill, New York.


298 Polynomial Approximation of Differential Equations<br />

gottlieb d.(1981), The stabil<strong>it</strong>y of pseudospectral Chebyshev methods, Math. Comp., 36, pp.107-118.<br />

gottlieb d., gunzburger m. & turkel e.(1982), On numerical boundary treatment for hyperbolic<br />

systems, SIAM J. Numer. Anal., 19, pp.671-697.<br />

gottlieb d., hussaini m.y. & orszag s.a.(1984), Theory and application of spectral methods, in Spectral<br />

Methods for Partial Differential Equations (R.G.Voigt, D.Gottlieb & M.Y.Hussaini Eds.),<br />

SIAM-CBMS, Philadelphia, pp.1-54.<br />

gottlieb d.& lustman.l.(1983), The spectrum of the Chebyshev collocation operator for the heat equation,<br />

SIAM J. Numer. Anal., 20, n.5, pp.909-921.<br />

gottlieb d., lustman.l.& tadmor e.(1987a), Stabil<strong>it</strong>y analysis of spectral methods for hyperbolic in<strong>it</strong>ialboundary<br />

value problems, SIAM J. Numer. Anal., 24, pp.241-256.<br />

gottlieb d., lustman.l.& tadmor e.(1987b), Convergence of spectral methods for hyperbolic in<strong>it</strong>ialboundary<br />

value problems, SIAM J. Numer. Anal., 24, pp.532-537.<br />

gottlieb d. & orszag s.a.(1977), Numerical Analysis of Spectral Methods, Theory and Applications,<br />

CBMS-NSF Regional Conference Series in Applied Mathematics, SIAM, Philadelphia.<br />

gottlieb d. & tadmor e.(1990), The CFL cond<strong>it</strong>ion for spectral approximations to hyperbolic in<strong>it</strong>ialboundary<br />

value problems, ICASE (Hampton VA) report n. 90-42.<br />

gottlieb d. & turkel e.(1985), Topics in spectral methods for time dependent problems, in Numerical<br />

Methods in Fluid Dynamics (F. Brezzi Ed.), Springer-Verlag, pp.115-155.<br />

greenspan d.(1961), Introduction to Partial Differential Equations, McGraw-Hill, New York.<br />

grosch c.e. & orszag s.a.(1977), Numerical solution of problems in unbounded regions: coordinate<br />

transformations, J. Comput. Phys., 25, pp.273-296.<br />

guelfand j.m., graev m.i. & vilenkin n.j.(1970), Les Distributions, 5 Volumes, Dunod, Paris.<br />

hackbusch w.(1985), Multi-Grid Methods and Applications, Spriger Series in Computational Mathematics,<br />

Springer-Verlag, Heidelberg.<br />

hageman l.a. & young d.m.(1981), Applied Iterative Methods, Academic Press, Orlando.<br />

haldenwang p., labrosse g., abboudi s. & deville m.(1984), Chebyshev 3-D spectral and 2-D<br />

pseudospectral solvers for the Helmholtz equation, J. Comput. Phys., 55, pp.115-128.<br />

halmos p.r.(1950), Measure Theory, D.Van Nostrand Company Inc., New York.<br />

hartman s. & mikusiński j.(1961), The Theory of Lebesgue Measure and Integration, Pergamon Press,<br />

Oxford.<br />

heinrichs w.(1988), Line relaxation for spectral multigrid methods, J. Comput. Phys., 77, n.1, pp.166-182.<br />

heinrichs w.(1989), Improved cond<strong>it</strong>ion number for spectral methods, Math. Comp., 53, n.187, pp.103-119.<br />

hildebrandt t.h.(1963), Theory of Integration, Academic Press, London.<br />

hill j.m. & dewynne j.n.(1987), Heat Conduction, Blackwell Scientific Publications, Oxford.<br />

hochstadt h.(1971), The functions of Mathematical Physics, Wiley-Interscience, New York.<br />

hochstadt h.(1973), Integral Equations, John Wiley & Sons, New York.<br />

hoffman k. & kunze r.(1971), Linear Algebra, Second ed<strong>it</strong>ion, Prentice-Hall, Englewood Cliffs NJ.<br />

ioakimidis n.i.(1981), On the weighted Galerkin method of numerical solution of Cauchy type singular<br />

integral equations, SIAM J. Numer. Anal., 18, n.6, pp.1120-1127.<br />

isaacson e. & keller h.b.(1966), Analysis of Numerical Methods, John Wiley & Sons, New York.<br />

jackson d.(1911), Über Genauigke<strong>it</strong> der Annäherung stetiger Funktionen durch ganze rationale Funktionen<br />

gegebenen Grades und trigonometrische Summen gegebener Ordnung, Dissertation, Göttingen.


References 299<br />

jackson d.(1930), The Theory of Approximation, American Mathematical Society, New York.<br />

jain m.k.(1984), Numerical Solution of Differential Equations, Second ed<strong>it</strong>ion, John Wiley & Sons, New<br />

York.<br />

kantorovich l.v. & krylov v.i.(1964), Approximate methods of Higher Analysis, Noordhoff, Groningen.<br />

karageorghis a. & phillips t.n.(1989), On the coefficients of differential expansions of ultraspherical<br />

polynomials, ICASE (Hampton VA) report n. 89-65.<br />

kavian o.(1990), Manuscript.<br />

kolmogorov a.n. & fomin s.v.(1961), Elements of the Theory of Functions and Functional Analysis,<br />

Graylock Press, Albany NY.<br />

kopriva d.a.(1986), A spectral multidomain method for the solution of hyperbolic systems, Applied Numer.<br />

Math., 2, pp.221-241.<br />

kosloff d. & tal-ezer h.(1989), Modified Chebyshev pseudospectral method w<strong>it</strong>h O(N −1 ) time step<br />

restriction, ICASE (Hampton VA) report n. 89-71.<br />

kreiss h.o. & lorenz j.(1989), In<strong>it</strong>ial-Boundary Value Problems and the Navier-Stokes Equation,<br />

Academic Press, London.<br />

krenk s.(1975), On quadrature formulas for singular integral equations of the first and the second kind,<br />

Quart. Appl. Math., 33, pp.225-232.<br />

krenk s.(1978), Quadrature formulae of closed type for solution of singular integral equations, J. Inst. Math.<br />

Appl., 22, pp.99-107.<br />

kufner a.(1980), Weighted Sobolev Spaces, Teubner Verlagsgesellschaft, Leipzig.<br />

ladyzhenskaya o.a.(1969), The Mathematical Theory of Viscous Incompressible Flow (translated from<br />

russian), Second ed<strong>it</strong>ion, Gordon and Breach Science Publishers, New York.<br />

lagerstrom p.a.(1988), Matched Asymptotic Expansions, Spinger-Verlag, New York.<br />

lanczos c.(1956), Applied Analysis, Prentice-Hall, Englewood Cliffs NJ.<br />

lapidus l. & seinfeld j.h.(1971), Numerical Solutions of Ordinary Differential Equations, Academic<br />

Press, New York.<br />

lax p.d. & milgram n.(1954), Parabolic equations, Contribution to the theory of partial differential<br />

equations, Ann. of Math. Stud., 33, pp.167-190.<br />

lions j.l. & magenes e.(1972), Non-Homogeneous Boundary Value Problems and Applications,<br />

Springer-Verlag, New York.<br />

lowenthal f.(1975), Linear Algebra w<strong>it</strong>h Linear Differential Equations, John Wiley & Sons, New York.<br />

luke y.l.(1969), The Special Functions and their Approximations, Academic Press, New York.<br />

lustman l.(1986), The time evolution of spectral discretization of hyperbolic equations, SIAM J. Numer.<br />

Anal., 23, pp.1193-1198.<br />

lynch r.e., rice j.r. & thomas d.h.(1964), Direct solution of partial differential equations by tensor<br />

product methods, Numer. Math., 6, pp.185-199.<br />

macaraeg m. & streett c.l.(1986), Improvements in spectral collocation through a multiple domain<br />

technique, Applied Numer. Math., 2, pp.95-108.<br />

maday y.(1990), Analysis of spectral projectors in one-dimensional domains, Math. Comp., 55, n.192,<br />

pp.537-562.<br />

maday y.(1991), Résultats d’approximation optimaux pur les opérateurs d’interpolation polynomiale, C.R.<br />

Acad. Sci. Paris, t.312, Série I, pp.705-710.


300 Polynomial Approximation of Differential Equations<br />

maday y., pernaud-thomas b. & vandeven h.(1985), Une réhabil<strong>it</strong>ation des méthodes spectrales de<br />

type Laguerre, Rech. Aérospat., 6, pp.353-375.<br />

maday y. & quarteroni a.(1981), Legendre and Chebyshev spectral approximations of Burger’s equation,<br />

Numer. Math., 37, pp.321-332.<br />

maday y. & quarteroni a.(1982), Approximation of Burger’s equation by pseudospectral methods,<br />

R.A.I.R.O. Anal. Numer., 16, pp.375-404.<br />

malik m.r., zang t.a. & hussaini m.y.(1985), A spectral collocation method for the Navier-Stokes<br />

equations, J. Comput. Phys., 61, pp.64-88.<br />

markoff a.(1889), Sur une question posée par Mendeleieff, Bulletin of the Academy of Sciences of St.<br />

Petersburg, 62, pp.1-24.<br />

markoff w.(1892), Über Polynome die in einen gegebenen Intervalle möglichst wenig von Null obweichen,<br />

Mathematische Annalen, Academy of Sciences of St. Petersburg, 77(1916), pp.213-258. From the<br />

original version in russian.<br />

mavriplis c.(1989), Laguerre polynomials for infin<strong>it</strong>e-domain spectral elements, J. Comput. Phys. ,80,<br />

pp.480-488.<br />

mercier b.(1982), Stabil<strong>it</strong>é et convergence des méthodes spectrales polynômiales. Application a l’équation<br />

d’advection, R.A.I.R.O., 16, n.1, pp.67-100.<br />

mercier b.(1989), An Introduction to the Numerical Analysis of Spectral Methods, Springer-Verlag,<br />

Berlin.<br />

monegato g.(1982), The numerical evaluation of one-dimensional Cauchy principal value integrals, Computing,<br />

29, pp.337-354.<br />

morchoisne y.(1979), Resolution of Navier-Stokes equations by a space-time pseudospectral method, Rech.<br />

Aérospat. 1979-5, pp.293-306.<br />

muñoz r.(1990), Theoretical analysis of some spectral multigrid method, in Spectral and High Order<br />

Methods for P.D.E. (C.Canuto & A.Quarteroni Eds.), North Holland, Amsterdam, pp.287-294.<br />

natanson i.p.(1965), Constructive Function Theory, Ungar, New York.<br />

něcas j.(1962), Sur une méthode pour resoudre les équations aux dérivées partielles du type elliptique voisine<br />

de la variationnelle, Ann. Scuola Norm. Sup. Pisa, 16, pp.305-326.<br />

nevai g.p.(1976), Lagrange interpolation at zeros of orthogonal polynomials, in Approximation Theory II<br />

(G.C.Lorenz, C.K.Chui & L.L.Schumaker Eds.), Academic Press, New York, pp.163-201.<br />

nochetto r.h.(1990), Numerical methods for free boundary problems, in Free Boundary Problems: Theory<br />

and Applications (K.H.Hoffman & J.Sprekels Eds.), P<strong>it</strong>man Research Notes in Mathematics.<br />

nussbaumer h.j.(1981), Fast Fourier Transform and Convolution Algor<strong>it</strong>hms, Springer Verlag, New<br />

York.<br />

okikiolu g.o.(1971), Aspects of the Theory of Bounded Integral Operators in L p -Spaces, Academic<br />

Press, London.<br />

orszag s.a.(1972), Comparison of pseudospectral and spectral approximations, Stud. Appl. Math., 51,<br />

pp.253-259.<br />

orszag s.a.(1980), Spectral methods for problems in complex geometries, J. Comput. Phys., 37, pp.70-92.<br />

orszag s.a.(1986), Fast eigenfunction transforms, Science and Computers Advances in Mathematics Supplementary<br />

Studies, 10, pp.23-30.<br />

osborn j.e.(1975), Spectral approximation for compact operators, Math. Comp., 29, n.131, pp.712-725.<br />

patera a.t.(1984), A spectral element method for fluid dynamics: laminar flow in a channel expansion, J.<br />

Comput. Phys., 54, pp.468-488.


References 301<br />

pavoni d.(1988), Single and multidomain Chebyshev collocation methods for the Korteweg-De Vries equation,<br />

Calcolo, 25, n.4, pp.311-346.<br />

peaceman d.w. & rachford h.h.Jr.(1955), The numerical solution of parabolic and elliptic differential<br />

equations, J. Soc. Indust. Appl. Math., 3, pp.28-41.<br />

pelz r.b.(1990), Pseudospectral methods on massively parallel computers, in Spectral and High Order<br />

Methods for P.D.E. (C.Canuto & A.Quarteroni Eds.), North Holland, Amsterdam, pp.493-503.<br />

pogorzelski w.(1966), Integral Equations and their Applications, Vol.1, Pergamon Press, Oxford.<br />

prandtl l.(1905), Über flüssigke<strong>it</strong>en bei sehr kleimer reibung, Proceedings of the Third International Mathematics<br />

Congress, Heidelberg, pp.484-491.<br />

quarteroni a.(1990), Domain decompos<strong>it</strong>ion methods for systems of conservation laws: spectral collocation<br />

approximations, SIAM J. Sci. Stat. Comput., 11, pp.1029-1052.<br />

quarteroni a. & sacchi-landriani g.(1988), Domain decompos<strong>it</strong>ion precond<strong>it</strong>ioners for the spectral<br />

collocation method, J. Sci. Comput., 3, n.1, pp.45-76.<br />

ralston a.(1965), A First Course in Numerical Analysis, Mc Graw-Hill, New York.<br />

rao m.m.(1987), Measure Theory and Integration, Wiley-Interscience, New York.<br />

rhee h.k., aris r. & amundson n.r.(1986), First-Order Partial Differential Equations. Volume I:<br />

Theory and Application of Single Equations, Prentice-Hall, Englewood Cliffs NJ.<br />

richardson l.f.(1910), The approximate ar<strong>it</strong>hmetical solution by fin<strong>it</strong>e differences of physical problems<br />

involving differential equations w<strong>it</strong>h an application to the stresses in a masonry dam, Phylos. Trans.<br />

Roy. Soc. London, A 210, pp.307-357.<br />

rivlin t.j.(1969), An Introduction to the Approximation of Functions, Blaisdell Publishing Company,<br />

Waltham MA.<br />

rivlin t.j.(1974), The Chebyshev polynomials, Wiley-Interscience, New York.<br />

rønquist e.m. & patera a.t.(1987), A Legendre spectral element method for the Stefan problem, Internat.<br />

J. Numer. Methods Engrg., 24, pp.2273-2299.<br />

rubinstein l.i.(1971), The Stefan Problem, Translation of Mathematical Monographs, 27, AMS, Providence<br />

RI.<br />

rudin w.(1966), Real and Complex Analysis, McGraw Hill, New York.<br />

schwarz h.a.(1890), Gesammelte Mathematische Abhandlungen, Springer, Berlin.<br />

schwartz l.(1966), Théorie des Distributions, Hermann, Paris.<br />

simmons g.f.(1972), Differential Equations w<strong>it</strong>h Applications and Historical Notes, McGraw-Hill, New<br />

York.<br />

singleton r.c.(1969), An algor<strong>it</strong>hm for computing the mixed radix fast Fourier transform, IEEE Trans.<br />

Audio Electroacoustic, AU15, pp.91-97.<br />

smoller j.(1983), Shock Waves and Reaction-Diffusion Equations, Springer-Verlag, New York.<br />

sneddon i.n.(1957), Elements of Partial Differential Equations, McGraw-Hill, New York.<br />

sobolev s.l.(1964), Partial Differential Equations of Mathematical Physics, Pergamon Press, Oxford.<br />

sod g.a.(1985), Numerical Methods in Fluid Dynamics, In<strong>it</strong>ial and In<strong>it</strong>ial Boundary-Value Problems,<br />

Cambridge Univers<strong>it</strong>y Press, Cambridge.<br />

solomonoff a. & turkel e.(1989), Global properties of pseudospectral methods, J. Comput. Phys., 81,<br />

pp.239-276.<br />

srivastava h.m. & manocha h.l.(1984), A Treatise on Generating Functions, Ellis Harwood, Chichester<br />

(England).


302 Polynomial Approximation of Differential Equations<br />

stekloff w.(1916), On approximate evaluation of defin<strong>it</strong>e integrals by means of formulas of mechanical<br />

quadratures. I: convergence of formulas of mechanical quadratures (in Russian), Bulletin de<br />

l’Académie Impériale des Sciences, Petrograd, 10, pp.169-186.<br />

strang g.(1972), Variational crimes in the fin<strong>it</strong>e element method, in Mathematical Fundation of the Fin<strong>it</strong>e<br />

Element Method w<strong>it</strong>h Applications to Partial Differential Equations (A.K. Aziz Ed.), Academic<br />

Press, New York, pp.689-710.<br />

strang g. & fix g.j.(1973), An Analysis of the Fin<strong>it</strong>e Element Method, Prentice-Hall, Englewood Cliffs<br />

NJ.<br />

streett c.l., zang t.a. & hussaini m.y.(1985), Spectral multigrid methods w<strong>it</strong>h applications to transonic<br />

potential flows, J. Comput. Phys., 57, pp.43-76.<br />

stroud a.h. & secrest d.h.(1966), Gaussian Quadrature Formulas Prentice-Hall, Englewood Cliffs NJ.<br />

swarztrauber p.n.(1986), Symmetric FFTs, Math. Comp., 47, pp.323-346.<br />

szegö g.(1939), Orthogonal polynomials, American Mathematical Society, New York.<br />

tadmor e.(1986), The exponential accuracy of Fourier and Chebyshev differencing methods, SIAM J. Numer.<br />

Anal., 23, n.1, pp.1-10.<br />

tal-ezer h.(1986a), Spectral methods in time for hyperbolic equations, SIAM J. Numer. Anal., 23, pp.11-<br />

26.<br />

tal-ezer h.(1986b), A pseudospectral Legendre method for hyperbolic equations w<strong>it</strong>h an improved stabil<strong>it</strong>y<br />

cond<strong>it</strong>ion, J. Comput. Phys., 67, pp.145-172.<br />

tal-ezer h.(1989), Spectral methods in time for parabolic equations, SIAM J. Numer. Anal., 26, n.1,<br />

pp.1-11.<br />

temam r.(1985), Navier-Stokes equations, Second ed<strong>it</strong>ion, North-Holland, Amsterdam.<br />

temperton c.(1983), Self-sorting mixed-radix fast Fourier transforms, J. Comput. Phys., 52, pp.1-23.<br />

temple g.(1971), The Structure of Lebesgue Integration Theory, Oxford Univers<strong>it</strong>y Press, London.<br />

timan a.f.(1963), Theory of Approximation of Functions of a Real Variable, Pergamon Press, Oxford.<br />

todd j.(1963), Introduction to the Constructive Theory of Functions, Birkhäuser Verlag, Basel.<br />

trefethen l.n.(1988), Lax-stabil<strong>it</strong>y vs. eigenvalue stabil<strong>it</strong>y of spectral methods, in Numerical Methods<br />

for Fluid Dynamics III (K.W.Morton & M.J.Baines Eds.), Clarendon Press, Oxford, pp.237-253.<br />

trefethen l.n. & trummer m.r.(1987), An instabil<strong>it</strong>y phenomenon in spectral methods, SIAM J. Numer.<br />

Anal., 24, n.5, pp.1008-1023.<br />

treves f.(1967), Topological Vector Spaces, Distributions and Kernels, Academic Press, New York.<br />

tricomi f.g.(1954), Funzioni Ipergeometriche Confluenti, Monografie Matematiche del C.N.R., Edizioni<br />

Cremonese, Rome.<br />

tricomi f.g.(1957), Integral Equations, Interscience Publisher, New York.<br />

triebel h.(1978), Interpolation Theory, Function Spaces, Differential Operators, North Holland, Amsterdam.<br />

trullinger s.e., zakharov v.e. & pokrovsky v.l. (Ed<strong>it</strong>ors) (1986), Sol<strong>it</strong>ons, Modern Problems in<br />

Condensed Matter Sciences, North Holland, Amsterdam.<br />

turkel e.(1980), Numerical methods for large-scale time-dependent partial differential equations, in Computational<br />

Fluid Dynamics, Vol.2, Hemisphere, Washington, pp.127-262.<br />

uspensky j.v.(1928), On the convergence of quadrature formulas related to an infin<strong>it</strong>e interval, Trans. Amer.<br />

Math. Soc., 30, pp.542-559.


References 303<br />

vainikko g.m.(1964), Asymptotic evaluations of the error of projection methods for the eigenvalue problem,<br />

USSR Computational Math. and Math. Phys., 4, pp.9-36.<br />

vainikko g.m.(1967), On the speed of convergence of approximate methods in the eigenvalue problem, USSR<br />

Computational Math. and Math. Phys., 7, pp.18-32.<br />

vandeven h.(1990), On the eigenvalues of second-order spectral differentiation operator, in Spectral and<br />

High Order Methods for P.D.E. (C.Canuto & A.Quarteroni Eds.) North Holland, Amsterdam,<br />

pp.313-318.<br />

varga r.s.(1962), Matrix Iterative Analysis, Prentice-Hall, Englewood Cliffs, NJ.<br />

vulikh b.z.(1963), Introduction to Functional Analysis for Scientists and Technologists, Pergamon Press,<br />

Oxford.<br />

watson g.(1966), A treatise on the Theory of Bessel Functions, Cambridge Univers<strong>it</strong>y Press, Cambridge.<br />

weideman j.a.c. & trefethen l.n.(1988), The eigenvalues of second-order spectral differentiation matrices,<br />

SIAM J. Numer. Anal., 25, n.6, pp.1279-1298.<br />

weinberger h.f.(1965), A first Course in Partial Differential Equations, John Wiley & Sons, New York.<br />

widder d.v.(1975), The Heat Equation, Academic Press, New York.<br />

wouk a.(1979), A Course of Applied Functional Analysis, John Wiley & Sons, New York.<br />

yosida k.(1960), Lectures on Differential and Integral Equations, Wiley-Interscience, New York.<br />

zampieri e.(1989), Iterative methods for domain decompos<strong>it</strong>ion using spectral tau and collocation approximation,<br />

Calcolo, 26, n.1, pp.61-91.<br />

zang t.a. & hussaini m.y.(1986), On spectral multigrid methods for time-dependent Navier-Stokes equations,<br />

Appl. Math. Comput., 19, pp.359-372.<br />

zang t.a., wong y.s. & hussaini m.y.(1982), Spectral multigrid methods for elliptic equations, J.<br />

Comput. Phys., 48, pp.485-501.<br />

zanolli p.(1987), Domain decompos<strong>it</strong>ion algor<strong>it</strong>hms for spectral methods, Calcolo, 24, pp.201-240.<br />

zygmund a.(1988), Trigonometric Series, Second ed<strong>it</strong>ion, Volumes I & II combined, Cambridge Univers<strong>it</strong>y<br />

Press, Cambridge.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!