13.07.2015 Views

Chapter I: Review of important results in quantum Mechanics - LSU

Chapter I: Review of important results in quantum Mechanics - LSU

Chapter I: Review of important results in quantum Mechanics - LSU

SHOW MORE
SHOW LESS

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

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

<strong>Chapter</strong> I: <strong>Review</strong> <strong>of</strong> <strong>important</strong> <strong>results</strong> <strong>in</strong><strong>quantum</strong> <strong>Mechanics</strong>1.1 Spectroscopy and <strong>quantum</strong> mechanicsSpectroscopy: an experimental subject concerned with the absorption, emission or scatter<strong>in</strong>g<strong>of</strong> electromagnetic radiation by atoms or molecules.Electromagnetic radiation covers a wide wavelength range, from radio waves to -rays, andthe atoms or molecules may be <strong>in</strong> the gas, liquid or solid phase or, <strong>of</strong> great importance <strong>in</strong>surface chemistry, adsorbed on a solid surface.Quantum mechanics: a theoretical subject relat<strong>in</strong>g to many aspects <strong>of</strong> chemistry and physics,but particularly to spectroscopy.Historical evolution:1665: Newton’s famous experiments on the dispersion <strong>of</strong> white light <strong>in</strong>to a range <strong>of</strong> coloursus<strong>in</strong>g a triangular glass prism.1860: Bunsen and Kirchh<strong>of</strong>f developed the prism spectroscope as an <strong>in</strong>tegrated unit for useas an analytical <strong>in</strong>strument. Early observations: emission spectra <strong>of</strong> various samples <strong>in</strong> aflame, the orig<strong>in</strong> <strong>of</strong> flame tests for various elements, and <strong>of</strong> the sun. visible spectrum <strong>of</strong>atomic hydrogen was observed both <strong>in</strong> the solar spectrum and <strong>in</strong> an electrical discharge <strong>in</strong>molecular hydrogen many years earlier.Late n<strong>in</strong>eteenth century: evidence accumulated that classical newtonian mechanics, whichwas completely successful on a macroscopic scale, was unsuccessful when applied toproblems on an atomic scale.1885: Balmer fitted the result<strong>in</strong>g series <strong>of</strong> l<strong>in</strong>es to a mathematical formula. He fitted thediscrete wavelengths <strong>of</strong> part <strong>of</strong> the emission spectrum <strong>of</strong> the hydrogen atom, now called theBalmer series and illustrated <strong>in</strong> Figure 1.1, to the empirical formula (1.1)1


where is a constant and In this figure, the wavenumber and thewavelength are used and are related by (<strong>in</strong> cm-1 ) (1.2)Us<strong>in</strong>g the relationship (<strong>in</strong> sec-1 ) (1.3)where is the frequency and the speed <strong>of</strong> light <strong>in</strong> a vacuum, Equation (1.1) becomes (1.4) <strong>in</strong> which is the Rydberg constant for hydrogen. The discrete rather than cont<strong>in</strong>uous nature<strong>of</strong> the spectrum is completely at variance with classical mechanics.1887: Hertz observes that when ultraviolet light falls on an alkali metal surface, electrons areejected from the surface only when the frequency <strong>of</strong> the radiation reaches a thresholdfrequency , which depends on the metal. As <strong>in</strong>cident frequency <strong>in</strong>creases, the k<strong>in</strong>etic energy<strong>of</strong> the ejected electrons (known as photoelectrons) <strong>in</strong>creases l<strong>in</strong>early with (Figure 1.2).Photoelectric effect was unexpla<strong>in</strong>ed by classical physics.Explanation <strong>of</strong> spectrum <strong>of</strong> hydrogen atom and <strong>of</strong> the photoelectric effect, together with otheranomalous observations such as the behaviour <strong>of</strong> the molar heat capacity <strong>of</strong> a solid attemperatures close to and the frequency distribution <strong>of</strong> black body radiation, orig<strong>in</strong>atedwith Planck.1900: Planck proposed that microscopic oscillators, <strong>of</strong> which a black body is made up, havean oscillation frequency related to the energy <strong>of</strong> the emitted radiation by (1.5)where is an <strong>in</strong>teger; is known as the Planck constant and its presently accepted value is (1.6)The energy is said to be quantized <strong>in</strong> discrete packets, or quanta, each <strong>of</strong> energy .Because is very small (it can be considered the energy etalon <strong>of</strong> the microscopic world),quantization <strong>of</strong> energy <strong>in</strong> macroscopic systems had escaped notice.2


1906: E<strong>in</strong>ste<strong>in</strong> applied this theory to the photoelectric effect and showed that (1.7)where = energy associated with a <strong>quantum</strong> (Lewis called it a photon <strong>in</strong> 1924) <strong>of</strong> the<strong>in</strong>cident radiation, = k<strong>in</strong>etic energy <strong>of</strong> the photoelectron ejected with velocity : =ionization energy <strong>of</strong> the metal surface (also called the work function for solids).1913: Bohr comb<strong>in</strong>ed classical and <strong>quantum</strong> mechanics to expla<strong>in</strong> observation <strong>of</strong> the Balmerseries but also <strong>of</strong> the Lyman, Paschen, Brackett, Pfund, etc., series <strong>of</strong> the hydrogen atomemission spectrum (Figure 1.1). He assumed empirically that the electron can move only <strong>in</strong>specific circular orbits around the nucleus and that the angular momentum for an angle <strong>of</strong>rotation is given by where and def<strong>in</strong>es the particular orbit. Energy is emitted or absorbed when theelectron moves from an orbit with higher to one <strong>of</strong> lower , or vice versa. The energy <strong>of</strong> the electron (derived by classical mechanics) is:(1.8) (1.9)where = is the reduced mass <strong>of</strong> the system <strong>of</strong> electron e and proton p;e = electronic charge; = permittivity <strong>of</strong> a vacuum. E n ‘s are the energy levels <strong>in</strong> Figure 1.1but = is plotted. Zero <strong>of</strong> energy corresponds to , i.e. atom is ionized. Energylevels discrete below but cont<strong>in</strong>uous above, s<strong>in</strong>ce electron can be ejected with anyamount <strong>of</strong> k<strong>in</strong>etic energy.When the electron transfers from, say, a lower to an upper orbit, energy requiredis (from equation (1.9)) or, s<strong>in</strong>ce (1.10) (1.11)from comparison with Equ. (1.4): and for the Balmer series.3


Similarly: , and 5 for Lyman, Paschen, Brackett and Pfund series, …However, it is <strong>important</strong> to realize that there is an <strong>in</strong>f<strong>in</strong>ite number <strong>of</strong> series, e.g. series withhigh are observed <strong>in</strong> the <strong>in</strong>terstellar medium (rich <strong>in</strong> H) by radioastronomy. For example,the transition at .The Rydberg constant from Equ. (1.11) is: (1.12)Planck’s <strong>quantum</strong> theory: very successful <strong>in</strong> expla<strong>in</strong><strong>in</strong>g the hydrogen atom spectrum, thewavelength distribution <strong>of</strong> black body radiation, the photoelectric effect and the low-temperature heat capacities <strong>of</strong> solids, but also gave rise to apparent anomalies: <strong>in</strong> thephotoelectric effect, photons are particles, while <strong>in</strong> <strong>in</strong>terference and diffraction, light is awave.1924: de Broglie proposes the wave-particle duality: (1.13)momentum (a particle property) is related to the wavelength (a wave property).1925: Davisson and Germer showed that electron can be wave-like by their diffraction atsurface <strong>of</strong> crystall<strong>in</strong>e nickel. Their experiment formed the basis <strong>of</strong> the LEED (low-energyelectron diffraction) technique for <strong>in</strong>vestigat<strong>in</strong>g structure near the surface <strong>of</strong> crystall<strong>in</strong>ematerials.1930: Further confirmation by Mark and Wierl showed that transmission <strong>of</strong> an electron beamthrough a gas target resulted <strong>in</strong> diffraction.Wave-particle duality rationalises why electron <strong>in</strong> H atom may be only <strong>in</strong> particular orbitswith angular momentum given by Equation (1.8). In wave picture, circumference <strong>of</strong> anorbit <strong>of</strong> radius r must conta<strong>in</strong> an <strong>in</strong>tegral number <strong>of</strong> wavelengths (1.14)where , 1, for a stand<strong>in</strong>g wave to be set up (see Figure 1.3(a) for , and4


Figure 1.3(b) shows how a travell<strong>in</strong>g wave <strong>results</strong> when is not an <strong>in</strong>teger: the wave<strong>in</strong>terferes with itself and is destroyed).However, electron <strong>in</strong> an orbit as a stand<strong>in</strong>g wave poses the <strong>important</strong> question <strong>of</strong> where theelectron, regarded as a particle, is. We shall consider the answer to this for the case <strong>of</strong> anelectron travell<strong>in</strong>g with constant velocity <strong>in</strong> a direction . The de Broglie picture <strong>of</strong> this is <strong>of</strong>a wave with a specific wavelength travell<strong>in</strong>g <strong>in</strong> direction as <strong>in</strong> Figure 1.4(a), and it is clearthat we cannot specify where the electron is. At the other extreme we can consider theelectron as a particle which can be observed as a sc<strong>in</strong>tillation on a phosphorescent screen.Figure 1.4(b) shows how, if there is a large number <strong>of</strong> waves <strong>of</strong> different wavelengths andamplitudes travell<strong>in</strong>g <strong>in</strong> the direction, they may re<strong>in</strong>force each other at a particular value <strong>of</strong>, say, and cancel each other elsewhere. This superposition at is called a wave packetand we can say the electron is behav<strong>in</strong>g as if it were a particle at .For Figure 1.4(a), momentum <strong>of</strong> electron is def<strong>in</strong>ed its position is completely uncerta<strong>in</strong>.In Figure 1.4(b), is certa<strong>in</strong> but the wavelength, and therefore , is uncerta<strong>in</strong>.1927: Heisenberg’s uncerta<strong>in</strong>ty pr<strong>in</strong>ciple: (1.15)With h / 2. In extreme wave picture, and , <strong>in</strong> extreme particlepicture, and . Other <strong>important</strong> form <strong>of</strong> the uncerta<strong>in</strong>ty pr<strong>in</strong>ciple: (1.16)(time , energy ), i.e. if energy <strong>of</strong> a state known exactly, and . Such a statedoes not change with time and is known as a stationary state.1.3 The Schröd<strong>in</strong>ger equation and some <strong>of</strong> its solutionsMany references given on QM <strong>in</strong> the bibliography, end <strong>of</strong> this chapter. Here, we briefly recallthe development <strong>of</strong> the Schröd<strong>in</strong>ger equation and some <strong>of</strong> its solutions that are vital to the<strong>in</strong>terpretation <strong>of</strong> atomic and molecular spectra.5


1.3.1 The Schröd<strong>in</strong>ger EquationJust as travell<strong>in</strong>g light wave can be represented by a function <strong>of</strong> its amplitude at a particularposition and time, so it was proposed that a wave function , a function <strong>of</strong> position andtime, describes the amplitude <strong>of</strong> an electron wave.1926: Born relates the wave and particle views by say<strong>in</strong>g that we should speak not <strong>of</strong> a particlebe<strong>in</strong>g at a particular po<strong>in</strong>t at a particular time but <strong>of</strong> the probability <strong>of</strong> f<strong>in</strong>d<strong>in</strong>g the particle there.This probability is , where is the complex conjugate <strong>of</strong> (obta<strong>in</strong>ed by replac<strong>in</strong>g all <strong>in</strong> by . It follows that:=. Reasonable also that: (1.17) (1.18)1928: Dirac shows that, when relativity is taken <strong>in</strong>to account, this is not quite true, but we shall notbe concerned with the effects <strong>of</strong> relativity.The form postulated for the wave function is (1.19)where is a constant and is the action, which is related to the k<strong>in</strong>etic energy and the potentialenergy by (1.20) the sum <strong>of</strong> the k<strong>in</strong>etic and potential energies (i.e. same as the hamiltonian <strong>in</strong> classicalmechanics). From Equations (1.19) and (1.20): (1.21)The form <strong>of</strong> the hamiltonian <strong>in</strong> <strong>quantum</strong> mechanics is obta<strong>in</strong>ed by replac<strong>in</strong>g the k<strong>in</strong>etic energy <strong>in</strong>Equation (1.20), giv<strong>in</strong>g (1.22)6


(Symbol , called ‘del’), <strong>in</strong> cartesian coord<strong>in</strong>ates , known as the laplacian, is: (1.23)From Equ. (1.21) and (1.22), time-dependent Schröd<strong>in</strong>ger equation: In conventional (frequency doma<strong>in</strong>) spectroscopy, time-<strong>in</strong>dependent part is ma<strong>in</strong>ly <strong>of</strong> concern.Consider (for ease <strong>of</strong> manipulation), wave travell<strong>in</strong>g <strong>in</strong> the direction, and assume that: can be factorized <strong>in</strong>to a giv<strong>in</strong>g (1.25)Where time-dependent part and time-<strong>in</strong>dependent part . Comb<strong>in</strong>ation <strong>of</strong> Equ. (1.24), fora one-dimensional system, and Equ. (1.25) gives 7 (1.26)S<strong>in</strong>ce the left-hand side (LHS) is function <strong>of</strong> x only, and the right-hand side (RHS) function <strong>of</strong> tonly, they must both be constant. S<strong>in</strong>ce they have the same dimensions as (i.e. energy) weput them equal to . For LHS: (1.27) one-dimensional time-<strong>in</strong>dependent Schröd<strong>in</strong>ger equation (also called the wave equation): (1.28)with H= hamiltonian (equ. 1.22, but for one dimension only). conta<strong>in</strong>s , i.e. it is an operator(e.g., is an operator which operates on to give ). In Equ. (1.28) operat<strong>in</strong>g on by givesthe result <strong>of</strong> multiplied by an energy . For example: (1.29)We are concerned with stationary states, known sometimes as eigenstates. The wave functions forthese states may be referred to as eigenfunctions and the associated energies as the eigenvalues.See bibliography for methods <strong>of</strong> solv<strong>in</strong>g the Schröd<strong>in</strong>ger equation for and for various systems.Here are some systems <strong>of</strong> <strong>in</strong>terest here.


1.3.2 The hydrogen atomThe H atom (a proton and one electron), occupies a very <strong>important</strong> position <strong>in</strong> the development <strong>of</strong><strong>quantum</strong> mechanics because the Schröd<strong>in</strong>ger equation may be solved exactly for this system. Alsotrue for hydrogen-like atomic ions , , , etc., and simple one-electron molecular ionssuch as .In QM picture <strong>of</strong> H atom, total energy is quantized and has exactly the same values as <strong>in</strong> Equ.(1.9). Angular momentum <strong>of</strong> the electron <strong>in</strong> a particular orbit (also called orbital) may also takeonly discrete value, is a vector and is def<strong>in</strong>ed, therefore, by its magnitude and direction. In theclassical picture <strong>of</strong> an electron circulat<strong>in</strong>g <strong>in</strong> an orbit <strong>in</strong> the direction shown <strong>in</strong> Figure 1.5 thedirection <strong>of</strong> the correspond<strong>in</strong>g vector, which is also shown, is given by the right-hand screw rule.In a magnetic field, can take only certa<strong>in</strong> orientations with respect to the magnetic field direction,so that the component <strong>of</strong> <strong>in</strong> that direction can take only certa<strong>in</strong>, discrete values. Thisphenomenon is referred to as space quantization. The effect is known as the Zeeman effect.For the hydrogen atom: (1.30)Second term on RHS is coulombic potential energy for the attraction between charges and a distance apart. Laplacian is here def<strong>in</strong>ed <strong>in</strong> spherical coord<strong>in</strong>ates as: (1.31)= distance <strong>of</strong> a po<strong>in</strong>t from the orig<strong>in</strong>; = co-latitude; = azimuth (Figure 1.6). The wavefunctions correspond<strong>in</strong>g to the hamiltonian <strong>of</strong> Equation (1.30) may be factorized: (1.32) are angular wave functions. They describe distribution <strong>of</strong> over the surface <strong>of</strong> a sphere <strong>of</strong>radius , i.e. they are spherical harmonics. Quantum number is same as <strong>in</strong> Bohrtheory, is the azimuthal <strong>quantum</strong> number associated with the discrete orbital angular momentumvalues, and is known as the magnetic <strong>quantum</strong> number which <strong>results</strong> from the spacequantization <strong>of</strong> the orbital angular momentum:(1.33) (1.34)8


<strong>in</strong> Equation (1.32) can be factorized further to give (1.35)The functions are the associated Legendre polynomials (table 1.1), and are <strong>in</strong>dependent <strong>of</strong> .They are therefore the same for all one-electron atoms.The solution <strong>of</strong> Schröd<strong>in</strong>ger equation for radial wave functions (they are functions only <strong>of</strong>), follows a well-known mathematical procedure to produce the solutions known as the associatedLaguerre functions, <strong>of</strong> which a few are given <strong>in</strong> Table 1.2. The radius <strong>of</strong> the Bohr orbit for isgiven by (1.36)For hydrogen, , and a useful quantity is related to by (1.37)Orbitals are labelled accord<strong>in</strong>g to values <strong>of</strong> and . The symbols s, p, d, f, g, ... <strong>in</strong>dicate values <strong>of</strong> . Thus we speak <strong>of</strong> 1s, 2s, 2p, 3s, 3p, 3d, etc. orbitals where the 1, 2, 3, etc. referto the value <strong>of</strong> .There are three useful ways <strong>of</strong> represent<strong>in</strong>g graphically:1. vs (or ), Figure 1.7(a). It can be seen that and are always positive but changes from positive to negative and, at one value <strong>of</strong> , is zero.2. vs (or ), Figure 1.7(b). = probability <strong>of</strong> f<strong>in</strong>d<strong>in</strong>g electron between and, i.e. this plot represents the radial probability distribution <strong>of</strong> the electron.3. vs (or ), Figure 1.7(c). = radial charge density, and is the probability<strong>of</strong> f<strong>in</strong>d<strong>in</strong>g the electron <strong>in</strong> a volume element consist<strong>in</strong>g <strong>of</strong> a th<strong>in</strong> spherical shell <strong>of</strong> thickness, radius , and volume .Diagrammatic representations <strong>of</strong> the functions (Equ. 1.35) cannot be made unless weconvert them from imag<strong>in</strong>ary <strong>in</strong>to real functions (Except when ).In the absence <strong>of</strong> an electric or magnetic field all the functions with are folddegenerate ( functions with the same energy, and each a possible values <strong>of</strong> .A property <strong>of</strong> degenerate functions: l<strong>in</strong>ear comb<strong>in</strong>ations <strong>of</strong> them are also solutions <strong>of</strong> theSchröd<strong>in</strong>ger equation. e.g., just as and are solutions <strong>of</strong> the Sch. Eq., so are9


- (1.38) From Equations (1.32), (1.35) and (1.38), together with the functions <strong>in</strong> Table 1.1, itfollows thats<strong>in</strong>ce (1.39) (1.40)Equs (1.39) become (1.41) (1.42)In addition, the third degenerate wave function is always real, and we label it ,where (1.43)All wave functions are always real and so are the , , etc. wave functions for However, for with , it is necessary to form l<strong>in</strong>ear comb<strong>in</strong>ations <strong>of</strong> theimag<strong>in</strong>ary wave functions and , or and , to obta<strong>in</strong> real functions. The orbital wave functions for any are dist<strong>in</strong>guished by subscripts , and , and and . There are seven orbitals for any but we shall not consider them here.Figure 1.8 shows the real wave functions for the 1s, 2p and 3d orbitals <strong>in</strong> the form <strong>of</strong>polar diagrams. For the case <strong>of</strong> the orbital, the wave function (equ. 1.43) is <strong>in</strong>dependent <strong>of</strong> and proportional to . The polar diagram consists <strong>of</strong> po<strong>in</strong>ts on a surface obta<strong>in</strong>ed by mark<strong>in</strong>g<strong>of</strong>f, on l<strong>in</strong>es drawn outwards from the nucleus <strong>in</strong> all directions, distances proportional to at a constant value <strong>of</strong> . The result<strong>in</strong>g surface consists <strong>of</strong> two touch<strong>in</strong>g spheres (Figure 1.8,10


also shows polar diagrams for all 1s, 2p and 3d orbitals). Figure 1.7 shows that as .Polar diagrams are drawn as surface boundaries with<strong>in</strong> which approx. 90% <strong>of</strong> the wavefunction resides.For all orbitals (except 1s) there are regions <strong>in</strong> space where because, and/or . Electron density is zero and we call them nodal surfaces or nodes. For example,the orbital has a nodal plane, while each <strong>of</strong> the 3d orbitals has two nodal planes. In general,there are such angular nodes where . The 2s orbital has one spherical nodal plane, orradial node, as Figure 1.7 shows. Generally, there are radial nodes for an orbital (or if we count the one at <strong>in</strong>f<strong>in</strong>ity).Quantum mechanical solution: same as <strong>in</strong> the Bohr theory (Equation (1.9)), it must do s<strong>in</strong>cethe Bohr theory agrees exactly with experiment, except for the f<strong>in</strong>e structure <strong>of</strong> the spectrum,which our present <strong>quantum</strong> mechanical treatment has not expla<strong>in</strong>ed either. The probabilitydistributions <strong>of</strong> figs 1.7 and 1.8, conta<strong>in</strong> function assymptotically go<strong>in</strong>g to zero and nodalsurfaces with zero probability <strong>of</strong> f<strong>in</strong>d<strong>in</strong>g the electron. This is far from the classical picture <strong>of</strong> theelectron orbit<strong>in</strong>g round the nucleus like the moon orbit<strong>in</strong>g round the earth.Unlike the total energy, the <strong>quantum</strong> mechanical value <strong>of</strong> the orbital angular momentum issignificantly different from that <strong>in</strong> the Bohr theory given (equ. 1.8). It is now given by (1.44)where .An effect <strong>of</strong> space quantization <strong>of</strong> orbital angular momentum may be observed if a magneticfield is <strong>in</strong>troduced along the axis. The orbital angular momentum vector , <strong>of</strong> magnitude ,may take up only certa<strong>in</strong> orientations such that the component along the axis is given by (1.45)where . This is illustrated <strong>in</strong> Figure 1.9 for an electron <strong>in</strong> a orbital .1.3.3 Electron sp<strong>in</strong> and nuclear sp<strong>in</strong> angular momentumThe electron and the nucleus can each sp<strong>in</strong> on its own axis, and each has an angular momentumassociated with sp<strong>in</strong>n<strong>in</strong>g. This is the orig<strong>in</strong> <strong>of</strong> the electron sp<strong>in</strong> and the nuclear sp<strong>in</strong>.Quantum mechanical treatment: magnitude <strong>of</strong> the angular momentum due to the ‘sp<strong>in</strong>’ <strong>of</strong> oneelectron is11


(1.46)where s =only. This result cannot be derived from the Schröd<strong>in</strong>ger equation but only fromDirac’s relativistic <strong>quantum</strong> mechanics. Space quantization <strong>of</strong> this angular momentum <strong>results</strong> <strong>in</strong>the componentwhere only (Figure 1.10). (1.47)Similarly, the magnitude <strong>of</strong> the angular momentum due to nuclear sp<strong>in</strong> is given by (1.48)The nuclear sp<strong>in</strong> <strong>quantum</strong> number may be zero, half-<strong>in</strong>teger or <strong>in</strong>teger, depend<strong>in</strong>g on the nucleusconcerned. Nuclei conta<strong>in</strong> protons and neutrons, each with . The way <strong>in</strong> which the sp<strong>in</strong>angular momenta <strong>of</strong> these particles couple together determ<strong>in</strong>es the value <strong>of</strong> , a few be<strong>in</strong>g given <strong>in</strong>Table 1.3. It is essential for nuclear magnetic resonance spectroscopy that for the nucleibe<strong>in</strong>g studied, so that, for example,nuclei cannot be used.1.3.4 The Born–Oppenheimer approximationHamiltonian H (eq. 1.20): In molecule, k<strong>in</strong>etic energy T consists <strong>of</strong> contributions T e and T n fromthe motions <strong>of</strong> the electrons and nuclei, respectively. Potential energy comprises two repulsiveterms, and (electron-electron and nucleus-nucleus <strong>in</strong>teractions), and , due to attractive<strong>in</strong>teractions between the electrons and nuclei, giv<strong>in</strong>g (1.49)For fixed nuclei , and , and there is a set <strong>of</strong> electronic wave functions which satisfy the Schröd<strong>in</strong>ger equation (1.50)where (1.51)S<strong>in</strong>ce depends on nuclear coord<strong>in</strong>ates, because <strong>of</strong> the term, so do and .1927: Born–Oppenheimer approximation: assumes nuclei move so slowly compared with electronsthat and <strong>in</strong>volve the nuclear coord<strong>in</strong>ates as parameters only. The result for a diatomicmolecule is that a curve <strong>of</strong> potential energy aga<strong>in</strong>st <strong>in</strong>ternuclear distance (or the displacement12


from equilibrium) can be drawn for a particular electronic state <strong>in</strong> which and are constant.Born–Oppenheimer approximation is valid because the electrons adjust <strong>in</strong>stantaneously to anynuclear motion: they are said to follow the nuclei. For this reason can be treated as part <strong>of</strong> thepotential field <strong>in</strong> which the nuclei move, so that (1.52)and the Schröd<strong>in</strong>ger equation for nuclear motion is (1.53)It follows BO approximation that total wave function can be factorized: (1.54)where the are electron coord<strong>in</strong>ates and is a function <strong>of</strong> nuclear coord<strong>in</strong>ates as well as . Itfollows from Equation (1.54) that (1.55)The wave function can be factorized further <strong>in</strong>to a vibrational part and a rotational part : (1.56)It follows that: (1.57)so thatand (1.58) (1.59)If any atoms have nuclear sp<strong>in</strong> this part <strong>of</strong> the total wave function can be factorized and the energytreated additively. It is for these reasons that we can treat electronic, vibrational, rotational andNMR spectroscopy separately.1.3.5 The harmonic oscillatorA good illustration for the vibration <strong>of</strong> a diatomic molecule is the ball-and-spr<strong>in</strong>g model. For smalldisplacements the stretch<strong>in</strong>g and compression <strong>of</strong> the bond, represented by the spr<strong>in</strong>g, obeysHooke’s law:13


(1.60)= potential energy, = force constant, the magnitude <strong>of</strong> which reflects the strength <strong>of</strong> the bond,= = displacement from the equilibrium bond length . Integration <strong>of</strong> this equation givesIn Figure 1.11, vs , illustrates the parabolic relationship.QM hamiltonian for one-dimensional harmonic oscillator: (1.61) (1.62)where = reduced mass <strong>of</strong> the nuclei. Schröd<strong>in</strong>ger equation (Equation 1.27) becomesfrom which it can be shown that v = 1,2,3,… and is the classical vibration frequency given by (1.63) (1.64) (1.65)Commonly, we use vibration wavenumber : (1.66)Equation (1.66) shows the vibrational levels to be equally spaced, by , and that the levelhas an energy , known as the zero-po<strong>in</strong>t energy. M<strong>in</strong>imum energy <strong>of</strong> molecule even at 0 °K,consequence <strong>of</strong> the uncerta<strong>in</strong>ty pr<strong>in</strong>ciple.Each po<strong>in</strong>t <strong>of</strong> <strong>in</strong>tersection <strong>of</strong> an energy level with the curve corresponds to a classical turn<strong>in</strong>g po<strong>in</strong>t<strong>of</strong> a vibration where the velocity <strong>of</strong> the nuclei is zero and all the energy is <strong>in</strong> the form <strong>of</strong> potentialenergy. This is <strong>in</strong> contrast to the mid-po<strong>in</strong>t <strong>of</strong> each energy level where all the energy is k<strong>in</strong>eticenergy.The wave functions from solutions <strong>of</strong> Equation (1.63) are (1.70)14


where the are known as Hermite polynomials and v (1.71)A few are given <strong>in</strong> Table 1.4 and some are plotted <strong>in</strong> Figure 1.13 where each energy levelcorresponds to . Important features <strong>of</strong> these wave functions:1. They penetrate regions outside the parabola, which would be forbidden <strong>in</strong> a classicalsystem.2. As <strong>in</strong>creases, two po<strong>in</strong>ts <strong>of</strong> maximum vibrational probability are classical turn<strong>in</strong>gpo<strong>in</strong>ts. This is illustrated for for which A and B are the classical turn<strong>in</strong>g po<strong>in</strong>ts,<strong>in</strong> contrast to the situation for , for which the maximum probability is at the midpo<strong>in</strong>t<strong>of</strong> the level.3. The wavelength <strong>of</strong> the ripples <strong>in</strong> <strong>in</strong>creases away from the classical turn<strong>in</strong>g po<strong>in</strong>ts.This is more apparent as <strong>in</strong>creases and is pronounced for .Further read<strong>in</strong>gAtk<strong>in</strong>s, P. W. (1991) Quanta: A Handbook <strong>of</strong> Concepts, Oxford University Press, Oxford.Atk<strong>in</strong>s, P. W. (1996) Molecular Quantum <strong>Mechanics</strong>, Oxford University Press, Oxford.Landau, L. D. and Lifshitz, E. M. (1959) Quantum <strong>Mechanics</strong>, Pergamon Press, Oxford.Paul<strong>in</strong>g, L. (1985) Introduction to Quantum <strong>Mechanics</strong>, Dover, New York.15


Table 1.3 Hermite polynomials for v = 0 to 516


Figure 1.1: Energy levels (vertical l<strong>in</strong>es) and observed transitions (horizontal l<strong>in</strong>es) <strong>of</strong> thehydrogen atom, <strong>in</strong>clud<strong>in</strong>g the Lyman, Balmer, Paschen, Brackett and Pfund series.Figure 1.2: Variation <strong>of</strong> k<strong>in</strong>etic energy <strong>of</strong> photoelectrons with the frequency, , <strong>of</strong> <strong>in</strong>cidentradiation.17


Figure 1.3: (a) A stand<strong>in</strong>g wave for an electron <strong>in</strong> an orbit with n = 6; (b) A travell<strong>in</strong>g wave,result<strong>in</strong>g when n is not an <strong>in</strong>tegerFigure 1.4: a) wave due to an electron travell<strong>in</strong>g at specific velocity <strong>in</strong> the x-direction: b)superposition <strong>of</strong> waves <strong>of</strong> different wavelengths <strong>in</strong>terfer<strong>in</strong>g constructively near x=0.Figure 1.5: Direction <strong>of</strong> the angular momentum p for an electron <strong>in</strong> an orbit.18


Figure 1.6: Spherical polar coord<strong>in</strong>ates r, and <strong>of</strong> a po<strong>in</strong>t PFigure 1.7: Plots <strong>of</strong>: a) radial wave function Rnl, (b) radial probability distribution function2Rnl ,19


(c) radial charge density function2 24rRnl, as a function <strong>of</strong> Figure 1.8: Polar diagrammes <strong>of</strong> the 1s, 2p and 3d atomic orbitals show<strong>in</strong>g the distribution <strong>of</strong> theangular wave functions20


Figure 1.9: Space quantization <strong>of</strong> orbital angular momentum for l=3Figure 1.10: Space quantization <strong>of</strong> electron sp<strong>in</strong> angular momentum21


Figure 1.11: Plot <strong>of</strong> potential energy, V(r) vs bond length r, for the harmonic oscillator model forvibration; r e is the equilibrium bond length. A few energy levels (for ) andthe correspond<strong>in</strong>g wave functions are shown; and are the classical turn<strong>in</strong>g po<strong>in</strong>ts on thewave function for .22

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!