molecular quantum mechanicsdar es salaam delhi hong kong istanbul karachi kolkata kuala lumpur...

588
MOLECULAR QUANTUM MECHANICS, FOURTH EDITION Peter Atkins Ronald Friedman OXFORD UNIVERSITY PRESS

Upload: others

Post on 29-May-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

  • MOLECULARQUANTUM

    MECHANICS,FOURTH EDITION

    Peter AtkinsRonald Friedman

    OXFORD UNIVERSITY PRESS

  • M O L E C U L A R Q U A N T U M M E C H A N I C S

  • This page intentionally left blank

  • F O U R T H E D I T I O N

    Peter Atkins

    University of Oxford

    Ronald Friedman

    Indiana Purdue Fort Wayne

    AC

    MOLECULAR

    QUANTUM

    MECHANICS

  • ACGreat Clarendon Street, Oxford OX2 6DP

    Oxford University Press is a department of the University of Oxford.It furthers the University’s objective of excellence in research, scholarship,

    and education by publishing worldwide in

    Oxford New York

    Auckland Bangkok Buenos Aires Cape Town ChennaiDar es Salaam Delhi Hong Kong Istanbul Karachi Kolkata

    Kuala Lumpur Madrid Melbourne Mexico City Mumbai NairobiSão Paulo Shanghai Taipei Tokyo Toronto

    Oxford is a registered trade mark of Oxford University Pressin the UK and in certain other countries

    Published in the United Statesby Oxford University Press Inc., New York

    # Peter Atkins and Ronald Friedman 2005

    The moral rights of the authors have been asserted.Database right Oxford University Press (maker)

    First published 2005

    All rights reserved. No part of this publication may be reproduced,stored in a retrieval system, or transmitted, in any form or by any means,

    without the prior permission in writing of Oxford University Press,or as expressly permitted by law, or under terms agreed with the appropriate

    reprographics rights organization. Enquiries concerning reproductionoutside the scope of the above should be sent to the Rights Department,

    Oxford University Press, at the address above

    You must not circulate this book in any other binding or coverand you must impose this same condition on any acquirer

    British Library Cataloguing in Publication Data

    Data available

    Library of Congress Cataloging in Publication Data

    Data available

    ISBN 0--19--927498--3

    10 9 8 7 6 5 4 3 2 1

    Typeset by Newgen Imaging Systems (P) Ltd., Chennai, IndiaPrinted in Great Britain

    on acid-free paper by Ashford Colour Press

  • Table of contents

    Preface xiii

    Introduction and orientation 1

    1 The foundations of quantum mechanics 9

    2 Linear motion and the harmonic oscillator 43

    3 Rotational motion and the hydrogen atom 71

    4 Angular momentum 98

    5 Group theory 122

    6 Techniques of approximation 168

    7 Atomic spectra and atomic structure 207

    8 An introduction to molecular structure 249

    9 The calculation of electronic structure 287

    10 Molecular rotations and vibrations 342

    11 Molecular electronic transitions 382

    12 The electric properties of molecules 407

    13 The magnetic properties of molecules 436

    14 Scattering theory 473

    Further information 513

    Further reading 553

    Appendix 1 Character tables and direct products 557

    Appendix 2 Vector coupling coefficients 562

    Answers to selected problems 563

    Index 565

  • This page intentionally left blank

  • Detailed Contents

    Introduction and orientation 1

    0.1 Black-body radiation 1

    0.2 Heat capacities 3

    0.3 The photoelectric and Compton effects 4

    0.4 Atomic spectra 5

    0.5 The duality of matter 6

    P R O B L E M S 8

    1 The foundations of quantum mechanics 9

    Operators in quantum mechanics 9

    1.1 Linear operators 10

    1.2 Eigenfunctions and eigenvalues 10

    1.3 Representations 12

    1.4 Commutation and non-commutation 13

    1.5 The construction of operators 14

    1.6 Integrals over operators 15

    1.7 Dirac bracket notation 16

    1.8 Hermitian operators 17

    The postulates of quantum mechanics 19

    1.9 States and wavefunctions 19

    1.10 The fundamental prescription 20

    1.11 The outcome of measurements 20

    1.12 The interpretation of the wavefunction 22

    1.13 The equation for the wavefunction 23

    1.14 The separation of the Schrödinger equation 23

    The specification and evolution of states 25

    1.15 Simultaneous observables 25

    1.16 The uncertainty principle 27

    1.17 Consequences of the uncertainty principle 29

    1.18 The uncertainty in energy and time 30

    1.19 Time-evolution and conservation laws 30

    Matrices in quantum mechanics 32

    1.20 Matrix elements 32

    1.21 The diagonalization of the hamiltonian 34

    The plausibility of the Schrödinger equation 36

    1.22 The propagation of light 36

    1.23 The propagation of particles 38

    1.24 The transition to quantum mechanics 39

    P R O B L E M S 40

    2 Linear motion and the harmonicoscillator 43

    The characteristics of acceptable wavefunctions 43

    Some general remarks on the Schrödinger equation 44

    2.1 The curvature of the wavefunction 45

    2.2 Qualitative solutions 45

    2.3 The emergence of quantization 46

    2.4 Penetration into non-classical regions 46

    Translational motion 47

    2.5 Energy and momentum 48

    2.6 The significance of the coefficients 48

    2.7 The flux density 49

    2.8 Wavepackets 50

    Penetration into and through barriers 51

    2.9 An infinitely thick potential wall 51

    2.10 A barrier of finite width 52

    2.11 The Eckart potential barrier 54

    Particle in a box 55

    2.12 The solutions 56

    2.13 Features of the solutions 57

    2.14 The two-dimensional square well 58

    2.15 Degeneracy 59

    The harmonic oscillator 60

    2.16 The solutions 61

    2.17 Properties of the solutions 63

    2.18 The classical limit 65

    Translation revisited: The scattering matrix 66

    P R O B L E M S 68

  • 3 Rotational motion and the hydrogen atom 71

    Particle on a ring 71

    3.1 The hamiltonian and the Schrödinger

    equation 71

    3.2 The angular momentum 73

    3.3 The shapes of the wavefunctions 74

    3.4 The classical limit 76

    Particle on a sphere 76

    3.5 The Schrödinger equation and

    its solution 76

    3.6 The angular momentum of the particle 79

    3.7 Properties of the solutions 81

    3.8 The rigid rotor 82

    Motion in a Coulombic field 84

    3.9 The Schrödinger equation for

    hydrogenic atoms 84

    3.10 The separation of the relative coordinates 85

    3.11 The radial Schrödinger equation 85

    3.12 Probabilities and the radial

    distribution function 90

    3.13 Atomic orbitals 91

    3.14 The degeneracy of hydrogenic atoms 94

    P R O B L E M S 96

    4 Angular momentum 98

    The angular momentum operators 98

    4.1 The operators and their commutation

    relations 99

    4.2 Angular momentum observables 101

    4.3 The shift operators 101

    The definition of the states 102

    4.4 The effect of the shift operators 102

    4.5 The eigenvalues of the angular momentum 104

    4.6 The matrix elements of the angular

    momentum 106

    4.7 The angular momentum eigenfunctions 108

    4.8 Spin 110

    The angular momenta of composite systems 112

    4.9 The specification of coupled states 112

    4.10 The permitted values of the total angular

    momentum 113

    4.11 The vector model of coupled angular

    momenta 115

    4.12 The relation between schemes 117

    4.13 The coupling of several angular momenta 119

    P R O B L E M S 120

    5 Group theory 122

    The symmetries of objects 122

    5.1 Symmetry operations and elements 123

    5.2 The classification of molecules 124

    The calculus of symmetry 129

    5.3 The definition of a group 129

    5.4 Group multiplication tables 130

    5.5 Matrix representations 131

    5.6 The properties of matrix representations 135

    5.7 The characters of representations 137

    5.8 Characters and classes 138

    5.9 Irreducible representations 139

    5.10 The great and little orthogonality theorems 142

    Reduced representations 145

    5.11 The reduction of representations 146

    5.12 Symmetry-adapted bases 147

    The symmetry properties of functions 151

    5.13 The transformation of p-orbitals 151

    5.14 The decomposition of direct-product bases 152

    5.15 Direct-product groups 155

    5.16 Vanishing integrals 157

    5.17 Symmetry and degeneracy 159

    The full rotation group 161

    5.18 The generators of rotations 161

    5.19 The representation of the full rotation group 162

    5.20 Coupled angular momenta 164

    Applications 165

    P R O B L E M S 166

    viii j CONTENTS

  • 6 Techniques of approximation 168

    Time-independent perturbation theory 168

    6.1 Perturbation of a two-level system 169

    6.2 Many-level systems 171

    6.3 The first-order correction to the energy 172

    6.4 The first-order correction to the wavefunction 174

    6.5 The second-order correction to the energy 175

    6.6 Comments on the perturbation expressions 176

    6.7 The closure approximation 178

    6.8 Perturbation theory for degenerate states 180

    Variation theory 183

    6.9 The Rayleigh ratio 183

    6.10 The Rayleigh–Ritz method 185

    The Hellmann–Feynman theorem 187

    Time-dependent perturbation theory 189

    6.11 The time-dependent behaviour of a

    two-level system 189

    6.12 The Rabi formula 192

    6.13 Many-level systems: the variation of constants 193

    6.14 The effect of a slowly switched constant

    perturbation 195

    6.15 The effect of an oscillating perturbation 197

    6.16 Transition rates to continuum states 199

    6.17 The Einstein transition probabilities 200

    6.18 Lifetime and energy uncertainty 203

    P R O B L E M S 204

    7 Atomic spectra and atomic structure 207

    The spectrum of atomic hydrogen 207

    7.1 The energies of the transitions 208

    7.2 Selection rules 209

    7.3 Orbital and spin magnetic moments 212

    7.4 Spin–orbit coupling 214

    7.5 The fine-structure of spectra 216

    7.6 Term symbols and spectral details 217

    7.7 The detailed spectrum of hydrogen 218

    The structure of helium 219

    7.8 The helium atom 219

    7.9 Excited states of helium 222

    7.10 The spectrum of helium 224

    7.11 The Pauli principle 225

    Many-electron atoms 229

    7.12 Penetration and shielding 229

    7.13 Periodicity 231

    7.14 Slater atomic orbitals 233

    7.15 Self-consistent fields 234

    7.16 Term symbols and transitions of

    many-electron atoms 236

    7.17 Hund’s rules and the relative energies of terms 239

    7.18 Alternative coupling schemes 240

    Atoms in external fields 242

    7.19 The normal Zeeman effect 242

    7.20 The anomalous Zeeman effect 243

    7.21 The Stark effect 245

    P R O B L E M S 246

    8 An introduction to molecular structure 249

    The Born–Oppenheimer approximation 249

    8.1 The formulation of the approximation 250

    8.2 An application: the hydrogen molecule–ion 251

    Molecular orbital theory 253

    8.3 Linear combinations of atomic orbitals 253

    8.4 The hydrogen molecule 258

    8.5 Configuration interaction 259

    8.6 Diatomic molecules 261

    8.7 Heteronuclear diatomic molecules 265

    Molecular orbital theory of polyatomic

    molecules 266

    8.8 Symmetry-adapted linear combinations 266

    8.9 Conjugated p-systems 269

    8.10 Ligand field theory 274

    8.11 Further aspects of ligand field theory 276

    The band theory of solids 278

    8.12 The tight-binding approximation 279

    8.13 The Kronig–Penney model 281

    8.14 Brillouin zones 284

    P R O B L E M S 285

    CONTENTS j ix

  • 9 The calculation of electronic structure 287

    The Hartree–Fock self-consistent field method 288

    9.1 The formulation of the approach 288

    9.2 The Hartree–Fock approach 289

    9.3 Restricted and unrestricted Hartree–Fock

    calculations 291

    9.4 The Roothaan equations 293

    9.5 The selection of basis sets 296

    9.6 Calculational accuracy and the basis set 301

    Electron correlation 302

    9.7 Configuration state functions 303

    9.8 Configuration interaction 303

    9.9 CI calculations 305

    9.10 Multiconfiguration and multireference methods 308

    9.11 Møller–Plesset many-body perturbation theory 310

    9.12 The coupled-cluster method 313

    Density functional theory 316

    9.13 Kohn–Sham orbitals and equations 317

    9.14 Exchange–correlation functionals 319

    Gradient methods and molecular properties 321

    9.15 Energy derivatives and the Hessian matrix 321

    9.16 Analytical derivatives and the coupled

    perturbed equations 322

    Semiempirical methods 325

    9.17 Conjugated p-electron systems 326

    9.18 Neglect of differential overlap 329

    Molecular mechanics 332

    9.19 Force fields 333

    9.20 Quantum mechanics–molecular mechanics 334

    Software packages for

    electronic structure calculations 336

    P R O B L E M S 339

    10 Molecular rotations and vibrations 342

    Spectroscopic transitions 342

    10.1 Absorption and emission 342

    10.2 Raman processes 344

    Molecular rotation 344

    10.3 Rotational energy levels 345

    10.4 Centrifugal distortion 349

    10.5 Pure rotational selection rules 349

    10.6 Rotational Raman selection rules 351

    10.7 Nuclear statistics 353

    The vibrations of diatomic molecules 357

    10.8 The vibrational energy levels of diatomic

    molecules 357

    10.9 Anharmonic oscillation 359

    10.10 Vibrational selection rules 360

    10.11 Vibration–rotation spectra of diatomic molecules 362

    10.12 Vibrational Raman transitions of diatomic

    molecules 364

    The vibrations of polyatomic molecules 365

    10.13 Normal modes 365

    10.14 Vibrational selection rules for polyatomic

    molecules 368

    10.15 Group theory and molecular vibrations 369

    10.16 The effects of anharmonicity 373

    10.17 Coriolis forces 376

    10.18 Inversion doubling 377

    Appendix 10.1 Centrifugal distortion 379

    P R O B L E M S 380

    11 Molecular electronic transitions 382

    The states of diatomic molecules 382

    11.1 The Hund coupling cases 382

    11.2 Decoupling and L-doubling 384

    11.3 Selection rules 386

    Vibronic transitions 386

    11.4 The Franck–Condon principle 386

    11.5 The rotational structure of vibronic transitions 389

    The electronic spectra of polyatomic molecules 390

    11.6 Symmetry considerations 391

    11.7 Chromophores 391

    11.8 Vibronically allowed transitions 393

    11.9 Singlet–triplet transitions 395

    The fate of excited species 396

    11.10 Non-radiative decay 396

    x j CONTENTS

  • 11.11 Radiative decay 397

    11.12 The conservation of orbital symmetry 399

    11.13 Electrocyclic reactions 399

    11.14 Cycloaddition reactions 401

    11.15 Photochemically induced electrocyclic reactions 403

    11.16 Photochemically induced cycloaddition reactions 404

    P R O B L E M S 406

    12 The electric properties of molecules 407

    The response to electric fields 407

    12.1 Molecular response parameters 407

    12.2 The static electric polarizability 409

    12.3 Polarizability and molecular properties 411

    12.4 Polarizabilities and molecular spectroscopy 413

    12.5 Polarizabilities and dispersion forces 414

    12.6 Retardation effects 418

    Bulk electrical properties 418

    12.7 The relative permittivity and the electric

    susceptibility 418

    12.8 Polar molecules 420

    12.9 Refractive index 422

    Optical activity 427

    12.10 Circular birefringence and optical rotation 427

    12.11 Magnetically induced polarization 429

    12.12 Rotational strength 431

    P R O B L E M S 434

    13 The magnetic properties of molecules 436

    The descriptions of magnetic fields 436

    13.1 The magnetic susceptibility 436

    13.2 Paramagnetism 437

    13.3 Vector functions 439

    13.4 Derivatives of vector functions 440

    13.5 The vector potential 441

    Magnetic perturbations 442

    13.6 The perturbation hamiltonian 442

    13.7 The magnetic susceptibility 444

    13.8 The current density 447

    13.9 The diamagnetic current density 450

    13.10 The paramagnetic current density 451

    Magnetic resonance parameters 452

    13.11 Shielding constants 452

    13.12 The diamagnetic contribution to shielding 456

    13.13 The paramagnetic contribution to shielding 458

    13.14 The g-value 459

    13.15 Spin–spin coupling 462

    13.16 Hyperfine interactions 463

    13.17 Nuclear spin–spin coupling 467

    P R O B L E M S 471

    14 Scattering theory 473

    The formulation of scattering events 473

    14.1 The scattering cross-section 473

    14.2 Stationary scattering states 475

    Partial-wave stationary scattering states 479

    14.3 Partial waves 479

    14.4 The partial-wave equation 480

    14.5 Free-particle radial wavefunctions and the

    scattering phase shift 481

    14.6 The JWKB approximation and phase shifts 484

    14.7 Phase shifts and the scattering matrix element 486

    14.8 Phase shifts and scattering cross-sections 488

    14.9 Scattering by a spherical square well 490

    14.10 Background and resonance phase shifts 492

    14.11 The Breit–Wigner formula 494

    14.12 Resonance contributions to the scattering

    matrix element 495

    Multichannel scattering 497

    14.13 Channels for scattering 497

    14.14 Multichannel stationary scattering states 498

    14.15 Inelastic collisions 498

    14.16 The S matrix and multichannel resonances 501

    The Green’s function 502

    14.17 The integral scattering equation and Green’s

    functions 502

    14.18 The Born approximation 504

    Appendix 14.1 The derivation of the Breit–Wigner

    formula 508

    Appendix 14.2 The rate constant for reactive

    scattering 509

    P R O B L E M S 510

    CONTENTS j xi

  • Further information 513

    Classical mechanics 513

    1 Action 513

    2 The canonical momentum 515

    3 The virial theorem 516

    4 Reduced mass 518

    Solutions of the Schrödinger equation 519

    5 The motion of wavepackets 519

    6 The harmonic oscillator: solution by

    factorization 521

    7 The harmonic oscillator: the standard solution 523

    8 The radial wave equation 525

    9 The angular wavefunction 526

    10 Molecular integrals 527

    11 The Hartree–Fock equations 528

    12 Green’s functions 532

    13 The unitarity of the S matrix 533

    Group theory and angular momentum 534

    14 The orthogonality of basis functions 534

    15 Vector coupling coefficients 535

    Spectroscopic properties 537

    16 Electric dipole transitions 537

    17 Oscillator strength 538

    18 Sum rules 540

    19 Normal modes: an example 541

    The electromagnetic field 543

    20 The Maxwell equations 543

    21 The dipolar vector potential 546

    Mathematical relations 547

    22 Vector properties 547

    23 Matrices 549

    Further reading 553

    Appendix 1 557

    Appendix 2 562

    Answers to selected problems 563

    Index 565

    xii j CONTENTS

  • PREFACE

    Many changes have occurred over the editions of this text but we have

    retained its essence throughout. Quantum mechanics is filled with abstract

    material that is both conceptually demanding and mathematically challen-

    ging: we try, wherever possible, to provide interpretations and visualizations

    alongside mathematical presentations.

    One major change since the third edition has been our response to concerns

    about the mathematical complexity of the material. We have not sacrificed

    the mathematical rigour of the previous edition but we have tried in

    numerous ways to make the mathematics more accessible. We have intro-

    duced short commentaries into the text to remind the reader of the mathe-

    matical fundamentals useful in derivations. We have included more worked

    examples to provide the reader with further opportunities to see formulae in

    action. We have added new problems for each chapter. We have expanded the

    discussion on numerous occasions within the body of the text to provide

    further clarification for or insight into mathematical results. We have set aside

    Proofs and Illustrations (brief examples) from the main body of the text so

    that readers may find key results more readily. Where the depth of pre-

    sentation started to seem too great in our judgement, we have sent material to

    the back of the chapter in the form of an Appendix or to the back of the book

    as a Further information section. Numerous equations are tabbed with wwwto signify that on the Website to accompany the text [www.oup.com/uk/

    booksites/chemistry/] there are opportunities to explore the equations by

    substituting numerical values for variables.

    We have added new material to a number of chapters, most notably the

    chapter on electronic structure techniques (Chapter 9) and the chapter on

    scattering theory (Chapter 14). These two chapters present material that is at

    the forefront of modern molecular quantum mechanics; significant advances

    have occurred in these two fields in the past decade and we have tried to

    capture their essence. Both chapters present topics where comprehension

    could be readily washed away by a deluge of algebra; therefore, we con-

    centrate on the highlights and provide interpretations and visualizations

    wherever possible.

    There are many organizational changes in the text, including the layout of

    chapters and the choice of words. As was the case for the third edition, the

    present edition is a rewrite of its predecessor. In the rewriting, we have aimed

    for clarity and precision.

    We have a deep sense of appreciation for many people who assisted us in

    this endeavour. We also wish to thank the numerous reviewers of the text-

    book at various stages of its development. In particular, we would like to

    thank

    Charles Trapp, University of Louisville, USA

    Ronald Duchovic, Indiana Purdue Fort Wayne, USA

  • Karl Jalkanen, Technical University of Denmark, Denmark

    Mark Child, University of Oxford, UK

    Ian Mills, University of Reading, UK

    David Clary, University of Oxford, UK

    Stephan Sauer, University of Copenhagen, Denmark

    Temer Ahmadi, Villanova University, USA

    Lutz Hecht, University of Glasgow, UK

    Scott Kirby, University of Missouri-Rolla, USA

    All these colleagues have made valuable suggestions about the content and

    organization of the book as well as pointing out errors best spotted in private.

    Many individuals (too numerous to name here) have offered advice over the

    years and we value and appreciate all their insights and advice. As always, our

    publishers have been very helpful and understanding.

    PWA, Oxford

    RSF, Indiana University Purdue University Fort Wayne

    June 2004

    xiv j PREFACE

  • There are two approaches to quantum mechanics. One is to follow the

    historical development of the theory from the first indications that the

    whole fabric of classical mechanics and electrodynamics should be held

    in doubt to the resolution of the problem in the work of Planck, Einstein,

    Heisenberg, Schrödinger, and Dirac. The other is to stand back at a point

    late in the development of the theory and to see its underlying theore-

    tical structure. The first is interesting and compelling because the theory

    is seen gradually emerging from confusion and dilemma. We see experi-

    ment and intuition jointly determining the form of the theory and, above

    all, we come to appreciate the need for a new theory of matter. The second,

    more formal approach is exciting and compelling in a different sense: there is

    logic and elegance in a scheme that starts from only a few postulates, yet

    reveals as their implications are unfolded, a rich, experimentally verifiable

    structure.

    This book takes that latter route through the subject. However, to set the

    scene we shall take a few moments to review the steps that led to the revo-

    lutions of the early twentieth century, when some of the most fundamental

    concepts of the nature of matter and its behaviour were overthrown and

    replaced by a puzzling but powerful new description.

    0.1 Black-body radiation

    In retrospect—and as will become clear—we can now see that theoretical

    physics hovered on the edge of formulating a quantum mechanical descrip-

    tion of matter as it was developed during the nineteenth century. However, it

    was a series of experimental observations that motivated the revolution. Of

    these observations, the most important historically was the study of black-

    body radiation, the radiation in thermal equilibrium with a body that absorbs

    and emits without favouring particular frequencies. A pinhole in an otherwise

    sealed container is a good approximation (Fig. 0.1).

    Two characteristics of the radiation had been identified by the end of the

    century and summarized in two laws. According to the Stefan–Boltzmann

    law, the excitance, M, the power emitted divided by the area of the emitting

    region, is proportional to the fourth power of the temperature:

    M ¼ sT4 ð0:1Þ

    Introduction and orientation

    0.1 Black-body radiation

    0.2 Heat capacities

    0.3 The photoelectric and

    Compton effects

    0.4 Atomic spectra

    0.5 The duality of matter

  • The Stefan–Boltzmann constant, s, is independent of the material from whichthe body is composed, and its modern value is 56.7 nW m�2 K�4. So, a region

    of area 1 cm2 of a black body at 1000 K radiates about 6 W if all frequencies

    are taken into account. Not all frequencies (or wavelengths, with l¼ c/n),though, are equally represented in the radiation, and the observed peak moves

    to shorter wavelengths as the temperature is raised. According to Wien’s

    displacement law,

    lmaxT ¼ constant ð0:2Þ

    with the constant equal to 2.9 mm K.

    One of the most challenging problems in physics at the end of the nine-

    teenth century was to explain these two laws. Lord Rayleigh, with minor help

    from James Jeans,1 brought his formidable experience of classical physics to

    bear on the problem, and formulated the theoretical Rayleigh–Jeans law for

    the energy density e(l), the energy divided by the volume, in the wavelengthrange l to lþ dl:

    deðlÞ ¼ rðlÞ dl rðlÞ ¼ 8pkTl4

    ð0:3Þ

    where k is Boltzmann’s constant (k¼ 1.381 � 10�23 J K�1). This formulasummarizes the failure of classical physics. It suggests that regardless of

    the temperature, there should be an infinite energy density at very short

    wavelengths. This absurd result was termed by Ehrenfest the ultraviolet

    catastrophe.

    At this point, Planck made his historic contribution. His suggestion was

    equivalent to proposing that an oscillation of the electromagnetic field of

    frequency n could be excited only in steps of energy of magnitude hn, whereh is a new fundamental constant of nature now known as Planck’s constant.

    According to this quantization of energy, the supposition that energy can be

    transferred only in discrete amounts, the oscillator can have the energies 0,

    hn, 2hn, . . . , and no other energy. Classical physics allowed a continuousvariation in energy, so even a very high frequency oscillator could be excited

    with a very small energy: that was the root of the ultraviolet catastrophe.

    Quantum theory is characterized by discreteness in energies (and, as we shall

    see, of certain other properties), and the need for a minimum excitation

    energy effectively switches off oscillators of very high frequency, and hence

    eliminates the ultraviolet catastrophe.

    When Planck implemented his suggestion, he derived what is now called

    the Planck distribution for the energy density of a black-body radiator:

    rðlÞ ¼ 8phcl5

    e�hc=lkT

    1� e�hc=lkT ð0:4Þ

    This expression, which is plotted in Fig. 0.2, avoids the ultraviolet cata-

    strophe, and fits the observed energy distribution extraordinarily well if we

    take h¼6.626� 10�34 J s. Just as the Rayleigh–Jeans law epitomizes thefailure of classical physics, the Planck distribution epitomizes the inception of

    . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

    1. ‘It seems to me,’ said Jeans, ‘that Lord Rayleigh has introduced an unnecessary factor 8 by

    counting negative as well as positive values of his integers.’ (Phil. Mag., 91, 10 (1905).)

    25

    20

    15

    10

    5

    00 0.5 1.0 1.5 2.0

    �kT hc/

    �/(8

    π(kT

    )5 /

    (hc)

    4 )

    Fig. 0.2 The Planck distribution.

    Pinhole

    Detectedradiation

    Containerat a temperature T

    Fig. 0.1 A black-body emitter can be

    simulated by a heated container witha pinhole in the wall. The

    electromagnetic radiation is reflected

    many times inside the container andreaches thermal equilibrium with the

    walls.

    2 j INTRODUCTION AND ORIENTATION

  • quantum theory. It began the new century as well as a new era, for it was

    published in 1900.

    0.2 Heat capacities

    In 1819, science had a deceptive simplicity. Dulong and Petit, for example,

    were able to propose their law that ‘the atoms of all simple bodies have

    exactly the same heat capacity’ of about 25 J K�1 mol�1 (in modern units).

    Dulong and Petit’s rather primitive observations, though, were done at room

    temperature, and it was unfortunate for them and for classical physics when

    measurements were extended to lower temperatures and to a wider range of

    materials. It was found that all elements had heat capacities lower than

    predicted by Dulong and Petit’s law and that the values tended towards zero

    as T ! 0.Dulong and Petit’s law was easy to explain in terms of classical physics by

    assuming that each atom acts as a classical oscillator in three dimensions. The

    calculation predicted that the molar isochoric (constant volume) heat capa-

    city, CV,m, of a monatomic solid should be equal to 3R¼24.94 J K�1 mol�1,where R is the gas constant (R¼NAk, with NA Avogadro’s constant). Thatthe heat capacities were smaller than predicted was a serious embarrassment.

    Einstein recognized the similarity between this problem and black-body

    radiation, for if each atomic oscillator required a certain minimum energy

    before it would actively oscillate and hence contribute to the heat capacity,

    then at low temperatures some would be inactive and the heat capacity would

    be smaller than expected. He applied Planck’s suggestion for electromagnetic

    oscillators to the material, atomic oscillators of the solid, and deduced the

    following expression:

    CV;mðTÞ ¼ 3RfEðTÞ fEðTÞ ¼yET e

    yE=2T

    1� eyE=T

    � �2ð0:5aÞ

    where the Einstein temperature, yE, is related to the frequency of atomicoscillators by yE¼hn/k. The function CV,m(T)/R is plotted in Fig. 0.3, andclosely reproduces the experimental curve. In fact, the fit is not particularly

    good at very low temperatures, but that can be traced to Einstein’s

    assumption that all the atoms oscillated with the same frequency. When this

    restriction was removed by Debye, he obtained

    CV;mðTÞ ¼ 3RfDðTÞ fDðTÞ ¼ 3T

    yD

    � �3Z yD=T0

    x4ex

    ðex � 1Þ2dx ð0:5bÞ

    where the Debye temperature, yD, is related to the maximum frequency of theoscillations that can be supported by the solid. This expression gives a very

    good fit with observation.

    The importance of Einstein’s contribution is that it complemented

    Planck’s. Planck had shown that the energy of radiation is quantized;

    3

    2

    1

    0

    Einstein

    Debye

    0 0.5 1 1.5 2T /�

    CV

    ,m/R

    Fig. 0.3 The Einstein and Debye

    molar heat capacities. The

    symbol y denotes the Einsteinand Debye temperatures,respectively. Close to T¼0 theDebye heat capacity is

    proportional to T3.

    0.2 HEAT CAPACITIES j 3

  • Einstein showed that matter is quantized too. Quantization appears to be

    universal. Neither was able to justify the form that quantization took (with

    oscillators excitable in steps of hn), but that is a problem we shall solve laterin the text.

    0.3 The photoelectric and Compton effects

    In those enormously productive months of 1905–6, when Einstein formu-

    lated not only his theory of heat capacities but also the special theory

    of relativity, he found time to make another fundamental contribution

    to modern physics. His achievement was to relate Planck’s quantum

    hypothesis to the phenomenon of the photoelectric effect, the emission of

    electrons from metals when they are exposed to ultraviolet radiation. The

    puzzling features of the effect were that the emission was instantaneous when

    the radiation was applied however low its intensity, but there was no emis-

    sion, whatever the intensity of the radiation, unless its frequency exceeded a

    threshold value typical of each element. It was also known that the kinetic

    energy of the ejected electrons varied linearly with the frequency of the

    incident radiation.

    Einstein pointed out that all the observations fell into place if the elec-

    tromagnetic field was quantized, and that it consisted of bundles of energy

    of magnitude hn. These bundles were later named photons by G.N. Lewis,and we shall use that term from now on. Einstein viewed the photoelectric

    effect as the outcome of a collision between an incoming projectile, a

    photon of energy hn, and an electron buried in the metal. This pictureaccounts for the instantaneous character of the effect, because even one

    photon can participate in one collision. It also accounted for the frequency

    threshold, because a minimum energy (which is normally denoted F andcalled the ‘work function’ for the metal, the analogue of the ionization

    energy of an atom) must be supplied in a collision before photoejection can

    occur; hence, only radiation for which hn>F can be successful. The lineardependence of the kinetic energy, EK, of the photoelectron on the frequency

    of the radiation is a simple consequence of the conservation of energy,

    which implies that

    EK ¼ hn� F ð0:6Þ

    If photons do have a particle-like character, then they should possess a

    linear momentum, p. The relativistic expression relating a particle’s energy to

    its mass and momentum is

    E2 ¼ m2c4 þ p2c2 ð0:7Þ

    where c is the speed of light. In the case of a photon, E¼ hn and m¼ 0, so

    p ¼ hnc¼ h

    lð0:8Þ

    4 j INTRODUCTION AND ORIENTATION

  • This linear momentum should be detectable if radiation falls on an electron,

    for a partial transfer of momentum during the collision should appear as a

    change in wavelength of the photons. In 1923, A.H. Compton performed the

    experiment with X-rays scattered from the electrons in a graphite target, and

    found the results fitted the following formula for the shift in wavelength,

    dl¼ lf� li, when the radiation was scattered through an angle y:

    dl ¼ 2lC sin2 12 y ð0:9Þ

    where lC¼ h/mec is called the Compton wavelength of the electron(lC¼2.426 pm). This formula is derived on the supposition that a photondoes indeed have a linear momentum h/l and that the scattering event is like acollision between two particles. There seems little doubt, therefore, that

    electromagnetic radiation has properties that classically would have been

    characteristic of particles.

    The photon hypothesis seems to be a denial of the extensive accumulation

    of data that apparently provided unequivocal support for the view that

    electromagnetic radiation is wave-like. By following the implications of

    experiments and quantum concepts, we have accounted quantitatively for

    observations for which classical physics could not supply even a qualitative

    explanation.

    0.4 Atomic spectra

    There was yet another body of data that classical physics could not elucidate

    before the introduction of quantum theory. This puzzle was the observation

    that the radiation emitted by atoms was not continuous but consisted of

    discrete frequencies, or spectral lines. The spectrum of atomic hydrogen had a

    very simple appearance, and by 1885 J. Balmer had already noticed that their

    wavenumbers, ~nn, where ~nn¼ n/c, fitted the expression

    ~nn ¼ RH1

    22� 1

    n2

    � �ð0:10Þ

    where RH has come to be known as the Rydberg constant for hydrogen

    (RH¼ 1.097�105 cm�1) and n¼3, 4, . . . . Rydberg’s name is commemoratedbecause he generalized this expression to accommodate all the transitions in

    atomic hydrogen. Even more generally, the Ritz combination principle states

    that the frequency of any spectral line could be expressed as the difference

    between two quantities, or terms:

    ~nn ¼ T1 � T2 ð0:11Þ

    This expression strongly suggests that the energy levels of atoms are confined

    to discrete values, because a transition from one term of energy hcT1 to

    another of energy hcT2 can be expected to release a photon of energy hc~nn, orhn, equal to the difference in energy between the two terms: this argument

    0.4 ATOMIC SPECTRA j 5

  • leads directly to the expression for the wavenumber of the spectroscopic

    transitions.

    But why should the energy of an atom be confined to discrete values? In

    classical physics, all energies are permissible. The first attempt to weld

    together Planck’s quantization hypothesis and a mechanical model of an atom

    was made by Niels Bohr in 1913. By arbitrarily assuming that the angular

    momentum of an electron around a central nucleus (the picture of an atom

    that had emerged from Rutherford’s experiments in 1910) was confined to

    certain values, he was able to deduce the following expression for the per-

    mitted energy levels of an electron in a hydrogen atom:

    En ¼ �me4

    8h2e20 1

    n2n ¼ 1, 2, . . . ð0:12Þ

    where 1/m¼1/meþ 1/mp and e0 is the vacuum permittivity, a fundamentalconstant. This formula marked the first appearance in quantum mechanics of

    a quantum number, n, which identifies the state of the system and is used to

    calculate its energy. Equation 0.12 is consistent with Balmer’s formula and

    accounted with high precision for all the transitions of hydrogen that were

    then known.

    Bohr’s achievement was the union of theories of radiation and models of

    mechanics. However, it was an arbitrary union, and we now know that it is

    conceptually untenable (for instance, it is based on the view that an electron

    travels in a circular path around the nucleus). Nevertheless, the fact that he

    was able to account quantitatively for the appearance of the spectrum of

    hydrogen indicated that quantum mechanics was central to any description of

    atomic phenomena and properties.

    0.5 The duality of matter

    The grand synthesis of these ideas and the demonstration of the deep links

    that exist between electromagnetic radiation and matter began with Louis de

    Broglie, who proposed on the basis of relativistic considerations that with any

    moving body there is ‘associated a wave’, and that the momentum of the body

    and the wavelength are related by the de Broglie relation:

    l ¼ hp

    ð0:13Þ

    We have seen this formula already (eqn 0.8), in connection with the prop-

    erties of photons. De Broglie proposed that it is universally applicable.

    The significance of the de Broglie relation is that it summarizes a fusion

    of opposites: the momentum is a property of particles; the wavelength is

    a property of waves. This duality, the possession of properties that in classical

    physics are characteristic of both particles and waves, is a persistent theme

    in the interpretation of quantum mechanics. It is probably best to regard

    the terms ‘wave’ and ‘particle’ as remnants of a language based on a false

    6 j INTRODUCTION AND ORIENTATION

  • (classical) model of the universe, and the term ‘duality’ as a late attempt to

    bring the language into line with a current (quantum mechanical) model.

    The experimental results that confirmed de Broglie’s conjecture are the

    observation of the diffraction of electrons by the ranks of atoms in a metal

    crystal acting as a diffraction grating. Davisson and Germer, who performed

    this experiment in 1925 using a crystal of nickel, found that the diffraction

    pattern was consistent with the electrons having a wavelength given by

    the de Broglie relation. Shortly afterwards, G.P. Thomson also succeeded

    in demonstrating the diffraction of electrons by thin films of celluloid

    and gold.2

    If electrons—if all particles—have wave-like character, then we should

    expect there to be observational consequences. In particular, just as a wave of

    definite wavelength cannot be localized at a point, we should not expect

    an electron in a state of definite linear momentum (and hence wavelength) to

    be localized at a single point. It was pursuit of this idea that led Werner

    Heisenberg to his celebrated uncertainty principle, that it is impossible to

    specify the location and linear momentum of a particle simultaneously with

    arbitrary precision. In other words, information about location is at the

    expense of information about momentum, and vice versa. This com-

    plementarity of certain pairs of observables, the mutual exclusion of the

    specification of one property by the specification of another, is also a major

    theme of quantum mechanics, and almost an icon of the difference between it

    and classical mechanics, in which the specification of exact trajectories was a

    central theme.

    The consummation of all this faltering progress came in 1926 when Werner

    Heisenberg and Erwin Schrödinger formulated their seemingly different but

    equally successful versions of quantum mechanics. These days, we step

    between the two formalisms as the fancy takes us, for they are mathematically

    equivalent, and each one has particular advantages in different types of cal-

    culation. Although Heisenberg’s formulation preceded Schrödinger’s by a few

    months, it seemed more abstract and was expressed in the then unfamiliar

    vocabulary of matrices. Still today it is more suited for the more formal

    manipulations and deductions of the theory, and in the following pages we

    shall employ it in that manner. Schrödinger’s formulation, which was in terms

    of functions and differential equations, was more familiar in style but still

    equally revolutionary in implication. It is more suited to elementary mani-

    pulations and to the calculation of numerical results, and we shall employ it in

    that manner.

    ‘Experiments’, said Planck, ‘are the only means of knowledge at our

    disposal. The rest is poetry, imagination.’ It is time for that imagination

    to unfold.

    . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

    2. It has been pointed out by M. Jammer that J.J. Thomson was awarded the Nobel Prize for

    showing that the electron is a particle, and G.P. Thomson, his son, was awarded the Prize for

    showing that the electron is a wave. (See The conceptual development of quantum mechanics,

    McGraw-Hill, New York (1966), p. 254.)

    0.5 THE DUALITY OF MATTER j 7

  • P R O B L E M S

    0.1 Calculate the size of the quanta involved in theexcitation of (a) an electronic motion of period 1.0 fs,(b) a molecular vibration of period 10 fs, and (c) a pendulumof period 1.0 s.

    0.2 Find the wavelength corresponding to the maximum inthe Planck distribution for a given temperature, and showthat the expression reduces to the Wien displacement law atshort wavelengths. Determine an expression for the constantin the law in terms of fundamental constants. (This constantis called the second radiation constant, c2.)

    0.3 Use the Planck distribution to confirm theStefan–Boltzmann law and to derive an expression forthe Stefan–Boltzmann constant s.

    0.4 The peak in the Sun’s emitted energy occurs at about480 nm. Estimate the temperature of its surface on the basisof it being regarded as a black-body emitter.

    0.5 Derive the Einstein formula for the heat capacity of acollection of harmonic oscillators. To do so, use thequantum mechanical result that the energy of a harmonicoscillator of force constant k and mass m is one of the values(vþ 12)hv, with v¼ (1/2p)(k/m)

    1/2 and v¼ 0, 1, 2, . . . . Hint.Calculate the mean energy, E, of a collection of oscillatorsby substituting these energies into the Boltzmanndistribution, and then evaluate C¼ dE/dT.

    0.6 Find the (a) low temperature, (b) high temperatureforms of the Einstein heat capacity function.

    0.7 Show that the Debye expression for the heat capacity isproportional to T3 as T! 0.

    0.8 Estimate the molar heat capacities of metallic sodium(yD¼ 150 K) and diamond (yD¼ 1860 K) at roomtemperature (300 K).

    0.9 Calculate the molar entropy of an Einstein solid atT¼ yE. Hint. The entropy is S ¼

    R T0 ðCV=TÞdT. Evaluate the

    integral numerically.

    0.10 How many photons would be emitted per second by asodium lamp rated at 100 W which radiated all its energywith 100 per cent efficiency as yellow light of wavelength589 nm?

    0.11 Calculate the speed of an electron emitted from a cleanpotassium surface (F¼ 2.3 eV) by light of wavelength (a)300 nm, (b) 600 nm.

    0.12 When light of wavelength 195 nm strikes a certain metalsurface, electrons are ejected with a speed of 1.23� 106 m s�1.Calculate the speed of electrons ejected from the same metalsurface by light of wavelength 255 nm.

    0.13 At what wavelength of incident radiation do therelativistic and non-relativistic expressions for the ejectionof electrons from potassium differ by 10 per cent? That is,find l such that the non-relativistic and relativistic linearmomenta of the photoelectron differ by 10 per cent. UseF¼ 2.3 eV.

    0.14 Deduce eqn 0.9 for the Compton effect on the basis ofthe conservation of energy and linear momentum. Hint. Usethe relativistic expressions. Initially the electron is at restwith energy mec

    2. When it is travelling with momentum p itsenergy is ðp2c2 þm2e c4Þ1/2. The photon, with initialmomentum h/li and energy hni, strikes the stationaryelectron, is deflected through an angle y, and emerges withmomentum h/lf and energy hnf. The electron is initiallystationary (p¼ 0) but moves off with an angle y 0 to theincident photon. Conserve energy and both components oflinear momentum. Eliminate y 0, then p, and so arrive at anexpression for dl.

    0.15 The first few lines of the visible (Balmer) series in thespectrum of atomic hydrogen lie at l/nm¼ 656.46, 486.27,434.17, 410.29, . . . . Find a value of RH, the Rydbergconstant for hydrogen. The ionization energy, I, is theminimum energy required to remove the electron. Find itfrom the data and express its value in electron volts. How isI related to RH? Hint. The ionization limit corresponds ton!1 for the final state of the electron.

    0.16 Calculate the de Broglie wavelength of (a) a mass of1.0 g travelling at 1.0 cm s�1, (b) the same at 95 per cent ofthe speed of light, (c) a hydrogen atom at room temperature(300 K); estimate the mean speed from the equipartitionprinciple, which implies that the mean kinetic energy of anatom is equal to 32kT, where k is Boltzmann’s constant, (d)an electron accelerated from rest through a potentialdifference of (i) 1.0 V, (ii) 10 kV. Hint. For the momentumin (b) use p¼mv/(l� v2/c2)1/2 and for the speed in (d) use12mev

    2¼ eV, where V is the potential difference.

    0.17 Derive eqn 0.12 for the permitted energy levels for theelectron in a hydrogen atom. To do so, use the following(incorrect) postulates of Bohr: (a) the electron moves in acircular orbit of radius r around the nucleus and (b) theangular momentum of the electron is an integral multiple of�h, that is mevr ¼ n�h. Hint. Mechanical stability of theorbital motion requires that the Coulombic force ofattraction between the electron and nucleus equals thecentrifugal force due to the circular motion. The energy ofthe electron is the sum of the kinetic energy and potential(Coulombic) energy. For simplicity, use me rather than thereduced mass m.

    8 j INTRODUCTION AND ORIENTATION

  • The whole of quantum mechanics can be expressed in terms of a small set

    of postulates. When their consequences are developed, they embrace the

    behaviour of all known forms of matter, including the molecules, atoms, and

    electrons that will be at the centre of our attention in this book. This chapter

    introduces the postulates and illustrates how they are used. The remaining

    chapters build on them, and show how to apply them to problems of chemical

    interest, such as atomic and molecular structure and the properties of mole-

    cules. We assume that you have already met the concepts of ‘hamiltonian’ and

    ‘wavefunction’ in an elementary introduction, and have seen the Schrödinger

    equation written in the form

    Hc ¼ EcThis chapter establishes the full significance of this equation, and provides

    a foundation for its application in the following chapters.

    Operators in quantum mechanics

    An observable is any dynamical variable that can be measured. The principal

    mathematical difference between classical mechanics and quantum mechan-

    ics is that whereas in the former physical observables are represented by

    functions (such as position as a function of time), in quantum mechanics they

    are represented by mathematical operators. An operator is a symbol for an

    instruction to carry out some action, an operation, on a function. In most of

    the examples we shall meet, the action will be nothing more complicated than

    multiplication or differentiation. Thus, one typical operation might be

    multiplication by x, which is represented by the operator x� . Anotheroperation might be differentiation with respect to x, represented by the

    operator d/dx. We shall represent operators by the symbol O (omega) ingeneral, but use A, B, . . . when we want to refer to a series of operators.

    We shall not in general distinguish between the observable and the operator

    that represents that observable; so the position of a particle along the x-axis

    will be denoted x and the corresponding operator will also be denoted x (with

    multiplication implied). We shall always make it clear whether we are

    referring to the observable or the operator.

    We shall need a number of concepts related to operators and functions

    on which they operate, and this first section introduces some of the more

    important features.

    The foundations of quantummechanics

    Operators in quantum mechanics

    1.1 Linear operators

    1.2 Eigenfunctions and eigenvalues

    1.3 Representations

    1.4 Commutation and

    non-commutation

    1.5 The construction of operators

    1.6 Integrals over operators

    1.7 Dirac bracket notation

    1.8 Hermitian operators

    The postulates of quantum

    mechanics

    1.9 States and wavefunctions

    1.10 The fundamental prescription

    1.11 The outcome of measurements

    1.12 The interpretation of the

    wavefunction

    1.13 The equation for the

    wavefunction

    1.14 The separation of the Schrödinger

    equation

    The specification and evolution of

    states

    1.15 Simultaneous observables

    1.16 The uncertainty principle

    1.17 Consequences of the uncertainty

    principle

    1.18 The uncertainty in energy and

    time

    1.19 Time-evolution and conservation

    laws

    Matrices in quantum mechanics

    1.20 Matrix elements

    1.21 The diagonalization of the

    hamiltonian

    The plausibility of the Schrödinger

    equation

    1.22 The propagation of light

    1.23 The propagation of particles

    1.24 The transition to quantum

    mechanics

    1

  • 1.1 Linear operators

    The operators we shall meet in quantum mechanics are all linear. A linear

    operator is one for which

    Oðaf þ bgÞ ¼ aOf þ bOg ð1:1Þwhere a and b are constants and f and g are functions. Multiplication is a

    linear operation; so is differentiation and integration. An example of a non-

    linear operation is that of taking the logarithm of a function, because it is not

    true, for example, that log 2x¼ 2 log x for all x.

    1.2 Eigenfunctions and eigenvalues

    In general, when an operator operates on a function, the outcome is another

    function. Differentiation of sin x, for instance, gives cos x. However, incertain cases, the outcome of an operation is the same function multiplied by

    a constant. Functions of this kind are called ‘eigenfunctions’ of the operator.

    More formally, a function f (which may be complex) is an eigenfunction of an

    operator O if it satisfies an equation of the form

    Of ¼ of ð1:2Þwhere o is a constant. Such an equation is called an eigenvalue equation. Thefunction eax is an eigenfunction of the operator d/dx because (d/dx)eax¼ aeax,which is a constant (a) multiplying the original function. In contrast, eax

    2is

    not an eigenfunction of d/dx, because (d/dx)eax2 ¼ 2axeax2 , which is a con-

    stant (2a) times a different function of x (the function xeax2). The constant o

    in an eigenvalue equation is called the eigenvalue of the operator O.

    Example 1.1 Determining if a function is an eigenfunction

    Is the function cos(3xþ 5) an eigenfunction of the operator d2/dx2 and, if so,what is the corresponding eigenvalue?

    Method. Perform the indicated operation on the given function and see if

    the function satisfies an eigenvalue equation. Use (d/dx)sin ax¼ a cos ax and(d/dx)cos ax¼�a sin ax.

    Answer. The operator operating on the function yields

    d2

    dx2cosð3xþ 5Þ ¼ d

    dx�3 sinð3xþ 5Þð Þ ¼ �9 cosð3xþ 5Þ

    and we see that the original function reappears multiplied by the eigen-

    value �9.

    Self-test 1.1. Is the function e3xþ 5 an eigenfunction of the operator d2/dx2

    and, if so, what is the corresponding eigenvalue?

    [Yes; 9]

    An important point is that a general function can be expanded in terms of

    all the eigenfunctions of an operator, a so-called complete set of functions.

    10 j 1 THE FOUNDATIONS OF QUANTUM MECHANICS

  • That is, if fn is an eigenfunction of an operator O with eigenvalue on (so Ofn¼on fn), then

    1 a general function g can be expressed as the linear combination

    g ¼X

    n

    cnfn ð1:3Þ

    where the cn are coefficients and the sum is over a complete set of functions.

    For instance, the straight line g¼ ax can be recreated over a certain range bysuperimposing an infinite number of sine functions, each of which is an

    eigenfunction of the operator d2/dx2. Alternatively, the same function may be

    constructed from an infinite number of exponential functions, which are

    eigenfunctions of d/dx. The advantage of expressing a general function as a

    linear combination of a set of eigenfunctions is that it allows us to deduce the

    effect of an operator on a function that is not one of its own eigenfunctions.

    Thus, the effect of O on g in eqn 1.3, using the property of linearity, is simply

    Og ¼ OX

    n

    cnfn ¼X

    n

    cnOfn ¼X

    n

    cnonfn

    A special case of these linear combinations is when we have a set of

    degenerate eigenfunctions, a set of functions with the same eigenvalue. Thus,

    suppose that f1, f2, . . . , fk are all eigenfunctions of the operator O, and thatthey all correspond to the same eigenvalue o:

    Ofn ¼ ofn with n ¼ 1, 2, . . . , k ð1:4ÞThen it is quite easy to show that any linear combination of the functions fnis also an eigenfunction of O with the same eigenvalue o. The proof is asfollows. For an arbitrary linear combination g of the degenerate set of

    functions, we can write

    Og ¼ OXkn¼1

    cnfn ¼Xkn¼1

    cnOfn ¼Xkn¼1

    cnofn ¼ oXkn¼1

    cnfn ¼ og

    This expression has the form of an eigenvalue equation (Og¼og).

    Example 1.2 Demonstrating that a linear combination of degenerate

    eigenfunctions is also an eigenfunction

    Show that any linear combination of the complex functions e2ix and e�2ix is an

    eigenfunction of the operator d2/dx2, where i¼ (�1)1/2.Method. Consider an arbitrary linear combination ae2ixþ be�2ix and see if thefunction satisfies an eigenvalue equation.

    Answer. First we demonstrate that e2ix and e�2ix are degenerate eigenfunctions.

    d2

    dx2e�2ix ¼ d

    dxð�2ie�2ixÞ ¼ �4e�2ix

    . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

    1. See P.M. Morse and H. Feschbach, Methods of theoretical physics, McGraw-Hill, New York

    (1953).

    1.2 EIGENFUNCTIONS AND EIGENVALUES j 11

  • where we have used i2¼�1. Both functions correspond to the same eigen-value, �4. Then we operate on a linear combination of the functions.

    d2

    dx2ðae2ix þ be�2ixÞ ¼ �4ðae2ix þ be�2ixÞ

    The linear combination satisfies the eigenvalue equation and has the same

    eigenvalue (�4) as do the two complex functions.

    Self-test 1.2. Show that any linear combination of the functions sin(3x) and

    cos(3x) is an eigenfunction of the operator d2/dx2.

    [Eigenvalue is �9]

    A further technical point is that from n basis functions it is possible to con-struct n linearly independent combinations. A set of functions g1, g2, . . . , gn is

    said to be linearly independent if we cannot find a set of constants c1, c2, . . . ,

    cn (other than the trivial set c1¼ c2¼ ¼0) for whichXi

    cigi ¼ 0

    A set of functions that is not linearly independent is said to be linearly

    dependent. From a set of n linearly independent functions, it is possible to

    construct an infinite number of sets of linearly independent combinations,

    but each set can have no more than n members. For example, from three

    2p-orbitals of an atom it is possible to form any number of sets of linearlyindependent combinations, but each set has no more than three members.

    1.3 Representations

    The remaining work of this section is to put forward some explicit forms of

    the operators we shall meet. Much of quantum mechanics can be developed in

    terms of an abstract set of operators, as we shall see later. However, it is often

    fruitful to adopt an explicit form for particular operators and to express them

    in terms of the mathematical operations of multiplication, differentiation,

    and so on. Different choices of the operators that correspond to a particular

    observable give rise to the different representations of quantum mechanics,

    because the explicit forms of the operators represent the abstract structure of

    the theory in terms of actual manipulations.

    One of the most common representations is the position representation,

    in which the position operator is represented by multiplication by x (orwhatever coordinate is specified) and the linear momentum parallel to x is

    represented by differentiation with respect to x. Explicitly:

    Position representation: x! x� px !�h

    i

    qqx

    ð1:5Þ

    where �h ¼ h=2p. Why the linear momentum should be represented in pre-cisely this manner will be explained in the following section. For the time

    being, it may be taken to be a basic postulate of quantum mechanics.

    An alternative choice of operators is the momentum representation, in

    which the linear momentum parallel to x is represented by the operation of

    12 j 1 THE FOUNDATIONS OF QUANTUM MECHANICS

  • multiplication by px and the position operator is represented by differentia-tion with respect to px. Explicitly:

    Momentum representation: x! � �hi

    qqpx

    px ! px� ð1:6Þ

    There are other representations. We shall normally use the position repres-

    entation when the adoption of a representation is appropriate, but we shall

    also see that many of the calculations in quantum mechanics can be done

    independently of a representation.

    1.4 Commutation and non-commutation

    An important feature of operators is that in general the outcome of successive

    operations (A followed by B, which is denoted BA, or B followed by A,denoted AB) depends on the order in which the operations are carried out.

    That is, in general BA 6¼AB. We say that, in general, operators do notcommute. For example, consider the operators x and px and a specific

    function x2. In the position representation, (xpx)x2¼x(2�h/i)x¼ (2�h/i)x2,

    whereas (pxx)x2¼ pxx3¼ (3�h/i)x2. The operators x and px do not commute.

    The quantity AB�BA is called the commutator of A and B and is denoted[A, B]:

    ½A, B ¼ AB� BA ð1:7Þ

    It is instructive to evaluate the commutator of the position and linear

    momentum operators in the two representations shown above; the procedure

    is illustrated in the following example.

    Example 1.3 The evaluation of a commutator

    Evaluate the commutator [x,px] in the position representation.

    Method. To evaluate the commutator [A,B] we need to remember that the

    operators operate on some function, which we shall write f. So, evaluate [A,B]f

    for an arbitrary function f, and then cancel f at the end of the calculation.

    Answer. Substitution of the explicit expressions for the operators into [x,px]

    proceeds as follows:

    ½x, pxf ¼ ðxpx � pxxÞf ¼ x��h

    i

    qfqx� �h

    i

    qðxf Þqx

    ¼ x� �hi

    qfqx� �h

    if � x� �h

    i

    qfqx¼ i�hf

    where we have used (1/i)¼�i. This derivation is true for any function f,so in terms of the operators themselves,

    ½x, px ¼ i�h

    The right-hand side should be interpreted as the operator ‘multiply by the

    constant i�h’.

    Self-test 1.3. Evaluate the same commutator in the momentum representation.

    [Same]

    1.4 COMMUTATION AND NON-COMMUTATION j 13

  • 1.5 The construction of operators

    Operators for other observables of interest can be constructed from the ope-

    rators for position and momentum. For example, the kinetic energy operator

    T can be constructed by noting that kinetic energy is related to linear

    momentum by T¼ p2/2m where m is the mass of the particle. It follows thatin one dimension and in the position representation

    T ¼ p2x

    2m¼ 1

    2m

    �h

    i

    d

    dx

    � �2¼ � �h

    2m

    d2

    dx2ð1:8Þ

    In three dimensions the operator in the position representation is

    T ¼ � �h2

    2m

    q2

    qx2þ q

    2

    qy2þ q

    2

    qz2

    ( )¼ � �h

    2

    2mr2 ð1:9Þ

    The operator r2, which is read ‘del squared’ and called the laplacian, is thesum of the three second derivatives.

    The operator for potential energy of a particle in one dimension, V(x), is

    multiplication by the function V(x) in the position representation. The same is

    true of the potential energy operator in three dimensions. For example, in the

    position representation the operator for the Coulomb potential energy of an

    electron (charge �e) in the field of a nucleus of atomic number Z is themultiplicative operator

    V ¼ � Ze2

    4pe0r� ð1:10Þ

    where r is the distance from the nucleus to the electron. It is usual to omit the

    multiplication sign from multiplicative operators, but it should not be for-

    gotten that such expressions are multiplications.

    The operator for the total energy of a system is called the hamiltonian

    operator and is denoted H:

    H ¼ T þ V ð1:11Þ

    The name commemorates W.R. Hamilton’s contribution to the formulation

    of classical mechanics in terms of what became known as a hamiltonian

    function. To write the explicit form of this operator we simply substitute the

    appropriate expressions for the kinetic and potential energy operators in the

    chosen representation. For example, the hamiltonian for a particle of mass m

    moving in one dimension is

    H ¼ � �h2

    2m

    d2

    dx2þ VðxÞ ð1:12Þ

    where V(x) is the operator for the potential energy. Similarly, the hamiltonian

    operator for an electron of mass me in a hydrogen atom is

    H ¼ � �h2

    2mer2 � e

    2

    4pe0rð1:13Þ

    Although eqn 1.9 has explicitly

    used Cartesian coordinates, the

    relation between the kinetic energy

    operator and the laplacian is true

    in any coordinate system; for

    example, spherical polar

    coordinates.

    14 j 1 THE FOUNDATIONS OF QUANTUM MECHANICS

  • The general prescription for constructing operators in the position repres-

    entation should be clear from these examples. In short:

    1. Write the classical expression for the observable in terms of position

    coordinates and the linear momentum.

    2. Replace x by multiplication by x, and replace px by (�h/i)q/qx (and likewisefor the other coordinates).

    1.6 Integrals over operators

    When we want to make contact between a calculation done using operators

    and the actual outcome of an experiment, we need to evaluate certain

    integrals. These integrals all have the form

    I ¼Z

    f �mOfn dt ð1:14Þ

    where f �m is the complex conjugate of fm. In this integral dt is the volumeelement. In one dimension, dt can be identified as dx; in three dimensions it isdxdydz. The integral is taken over the entire space available to the system,

    which is typically from x¼�1 to x¼ þ1 (and similarly for the othercoordinates). A glance at the later pages of this book will show that many

    molecular properties are expressed as combinations of integrals of this form

    (often in a notation which will be explained later). Certain special cases of this

    type of integral have special names, and we shall introduce them here.

    When the operator O in eqn 1.14 is simply multiplication by 1, the integralis called an overlap integral and commonly denoted S:

    S ¼Z

    f �mfn dt ð1:15Þ

    It is helpful to regard S as a measure of the similarity of two functions: whenS¼0, the functions are classified as orthogonal, rather like two perpendicularvectors. When S is close to 1, the two functions are almost identical. The

    recognition of mutually orthogonal functions often helps to reduce the

    amount of calculation considerably, and rules will emerge in later sections

    and chapters.

    The normalization integral is the special case of eqn 1.15 for m¼n.A function fm is said to be normalized (strictly, normalized to 1) ifZ

    f �mfm dt ¼ 1 ð1:16Þ

    It is almost always easy to ensure that a function is normalized by multiplying

    it by an appropriate numerical factor, which is called a normalization factor,

    typically denoted N and taken to be real so that N� ¼N. The procedure isillustrated in the following example.

    Example 1.4 How to normalize a function

    A certain function f is sin(px/L) between x¼ 0 and x¼L and is zero elsewhere.Find the normalized form of the function.

    The complex conjugate of

    a complex number z¼ aþ ibis z� ¼ a� ib. Complexconjugation amounts to

    everywhere replacing i by � i.The square modulus jzj2 is given byzz� ¼ a2þ b2 since jij2¼ 1.

    1.6 INTEGRALS OVER OPERATORS j 15

  • Method. We need to find the (real) factor N such that N sin(px/L) is norm-alized to 1. To find N we substitute this expression into eqn 1.16, evaluate the

    integral, and select N to ensure normalization. Note that ‘all space’ extends

    from x¼ 0 to x¼L.

    Answer. The necessary integration isZ

    f �f dt ¼Z L

    0

    N2 sin2ðpx=LÞdx ¼ 12LN2

    where we have usedR

    sin2ax dx¼ (x/2)(sin 2ax)/4aþ constant. For thisintegral to be equal to 1, we require N¼ (2/L)1/2. The normalized function istherefore

    f ¼ 2L

    � �1=2sinðpx=LÞ

    Comment. We shall see later that this function describes the distribution of a

    particle in a square well, and we shall need its normalized form there.

    Self-test 1.4. Normalize the function f¼ eif, where f ranges from 0 to 2p.[N¼ 1/(2p)1/2]

    A set of functions fn that are (a) normalized and (b) mutually orthogonal

    are said to satisfy the orthonormality condition:Zf �mfn dt ¼ dmn ð1:17Þ

    In this expression, dmn denotes the Kronecker delta, which is 1 when m¼nand 0 otherwise.

    1.7 Dirac bracket notation

    With eqn 1.14 we are on the edge of getting lost in a complicated notation. The

    appearance of many quantum mechanical expressions is greatly simplified by

    adopting the Dirac bracket notation in which integrals are written as follows:

    hmjOjni ¼Z

    f �mOfn dt ð1:18Þ

    The symbol jni is called a ket, and denotes the state described by the functionfn. Similarly, the symbol hnj is called a bra, and denotes the complex conjugateof the function, f �n . When a bra and ket are strung together with an operator

    between them, as in the bracket hmjOjni, the integral in eqn 1.18 is to beunderstood. When the operator is simply multiplication by 1, the 1 is omitted

    and we use the convention

    hmjni ¼Z

    f �mfn dt ð1:19Þ

    This notation is very elegant. For example, the normalization integral

    becomes hnjni¼1 and the orthogonality condition becomes hmjni¼0for m 6¼n. The combined orthonormality condition (eqn 1.17) is thenhmjni ¼ dmn ð1:20Þ

    16 j 1 THE FOUNDATIONS OF QUANTUM MECHANICS

  • A final point is that, as can readily be deduced from the definition of a Dirac

    bracket,

    hmjni ¼ hnjmi�

    1.8 Hermitian operators

    An operator is hermitian if it satisfies the following relation:Zf �mOfn dt ¼

    �Zf �nOfmdt

    �ð1:21aÞ

    for any two functions fm and fn. An alternative version of this definition isZf �mOfn dt ¼

    ZðOfmÞ�fn dt ð1:21bÞ

    This expression is obtained by taking the complex conjugate of each term on

    the right-hand side of eqn 1.21a. In terms of the Dirac notation, the definition

    of hermiticity is

    hmjOjni ¼ hnjOjmi� ð1:22Þ

    Example 1.5 How to confirm the hermiticity of operators

    Show that the position and momentum operators in the position representa-

    tion are hermitian.

    Method. We need to show that the operators satisfy eqn 1.21a. In some cases

    (the position operator, for instance), the hermiticity is obvious as soon as the

    integral is written down. When a differential operator is used, it may be

    necessary to use integration by parts at some stage in the argument to transfer

    the differentiation from one function to another:Zu dv ¼ uv�

    Zv du

    Answer. That the position operator is hermitian is obvious from inspection:Zf �mxfn dt ¼

    Zfnxf

    �m dt ¼

    Zf �n xfm dt

    � �

    We have used the facts that (f�)� ¼ f and x is real. The demonstration of thehermiticity of px, a differential operator in the position representation,

    involves an integration by parts:

    Zf �mpxfn dx ¼

    Zf �m

    �h

    i

    d

    dxfn dx ¼

    �h

    i

    Zf �m dfn

    ¼ �hi

    �f �mfn �

    Zfndf

    �m

    x¼1

    x¼�1

    ¼ �hi

    f �mfnjx¼1x¼�1 �Z 1�1

    fnd

    dxf �m dx

    1.8 HERMITIAN OPERATORS j 17

  • The first term on the right is zero (because when jxj is infinite, a normalizablefunction must be vanishingly small; see Section 1.12). Therefore,

    Zf �mpxfn dx ¼ �

    �h

    i

    Zfn

    d

    dxf �m dx

    ¼Z

    f �n�h

    i

    d

    dxfm dx

    � �¼

    Zf �n pxfm dx

    � �

    Hence, the operator is hermitian.

    Self-test 1.5. Show that the two operators are hermitian in the momentum

    representation.

    As we shall now see, the property of hermiticity has far-reaching impli-

    cations. First, we shall establish the following property:

    Property 1. The eigenvalues of hermitian operators are real.

    Proof 1.1 The reality of eigenvalues

    Consider the eigenvalue equation

    Ojoi ¼ ojoi

    The ket joi denotes an eigenstate of the operator O in the sense that thecorresponding function fo is an eigenfunction of the operator O and we arelabelling the eigenstates with the eigenvalue o of the operator O. It is oftenconvenient to use the eigenvalues as labels in this way. Multiplication from the

    left by hoj results in the equation

    hojOjoi ¼ ohojoi ¼ o

    taking joi to be normalized. Now take the complex conjugate of both sides:

    hojOjoi� ¼ o�

    However, by hermiticity, hojOjoi� ¼ hojOjoi. Therefore, it follows thato¼o�, which implies that the eigenvalue o is real.

    The second property we shall prove is as follows:

    Property 2. Eigenfunctions corresponding to different eigenvalues of an

    hermitian operator are orthogonal.

    That is, if we have two eigenfunctions of an hermitian operator O witheigenvalues o and o 0, with o 6¼o 0, then hojo 0i ¼ 0. For example, it follows atonce that all the eigenfunctions of a harmonic oscillator (Section 2.16) are

    mutually orthogonal, for as we shall see each one corresponds to a different

    energy (the eigenvalue of the hamiltonian, an hermitian operator).

    18 j 1 THE FOUNDATIONS OF QUANTUM MECHANICS

  • Proof 1.2 The orthogonality of eigenstates

    Suppose we have two eigenstates joi and jo 0i that satisfy the followingrelations:

    Ojoi ¼ ojoi and Ojo0i ¼ o0jo0iThen multiplication of the first relation by ho 0j and the second by hoj givesho0jOjoi ¼ oho0joi and hojOjo0i ¼ o0hojo0i

    Now take the complex conjugate of the second relation and subtract it from

    the first while using Property 1 (o 0� ¼o 0):ho0jOjoi � hojOjo0i� ¼ oho0joi � o0hojo0i�

    Because O is hermitian, the left-hand side of this expression is zero; so (notingthat o 0 is real and using hojo 0i� ¼ ho 0joi as explained earlier) we arrive atðo� o0Þho0joi ¼ 0

    However, because the two eigenvalues are different, the only way of satisfying

    this relation is for ho 0joi¼ 0, as was to be proved.

    The postulates of quantum mechanics

    Now we turn to an application of the preceding material, and move into the

    foundations of quantum mechanics. The postulates we use as a basis for

    quantum mechanics are by no means the most subtle that have been devised,

    but they are strong enough for what we have to do.

    1.9 States and wavefunctions

    The first postulate concerns the information we can know about a state:

    Postulate 1. The state of a system is fully described by a function C(r1,r2, . . . , t).

    In this statement, r1, r2, . . . are the spatial coordinates of particles 1, 2, . . .

    that constitute the system and t is the time. The function C (uppercase psi)plays a central role in quantum mechanics, and is called the wavefunction of

    the system (more specifically, the time-dependent wavefunction). When we

    are not interested in how the system changes in time we shall denote the

    wavefunction by a lowercase psi as c(r1, r2, . . . ) and refer to it as the time-independent wavefunction. The state of the system may also depend on some

    internal variable of the particles (their spin states); we ignore that for now

    and return to it later. By ‘describe’ we mean that the wavefunction

    contains information about all the properties of the system that are open to

    experimental determination.

    We shall see that the wavefunction of a system will be specified by a set of

    labels called quantum numbers, and may then be written ca,b, . . . , wherea, b, . . . are the quantum numbers. The values of these quantum numbers

    specify the wavefunction and thus allow the values of various physical

    1.9 STATES AND WAVEFUNCTIONS j 19

  • observables to be calculated. It is often convenient to refer to the state of

    the system without referring to the corresponding wavefunction; the state is

    specified by listing the values of the quantum numbers that define it.

    1.10 The fundamental prescription

    The next postulate concerns the selection of operators:

    Postulate 2. Observables are represented by hermitian operators chosen to

    satisfy the commutation relations

    ½q, pq0 ¼ i�hdqq0 ½q, q0 ¼ 0 ½pq, pq0 ¼ 0where q and q 0 each denote one of the coordinates x, y, z and pq and pq 0 the

    corresponding linear momenta.

    The requirement that the operators are hermitian ensures that the observables

    have real values (see below). Each commutation relation is a basic, unpro-

    vable, and underivable postulate. Postulate 2 is the basis of the selection of

    the form of the operators in the position and momentum representations for

    all observables that depend on the position and the momentum.2 Thus, if we

    define the position representation as the representation in which the position

    operator is multiplication by the position coordinate, then as we saw in

    Example 1.3, it follows that the momentum operator must involve differ-

    entiation with respect to x, as specified earlier. Similarly, if the momentum

    representation is defined as the representation in which the linear momentum

    is represented by multiplication, then the form of the position operator is

    fixed as a derivative with respect to the linear momentum. The coordinates

    x, y, and z commute with each other as do the linear momenta px, py, and pz.

    1.11 The outcome of measurements

    The next postulate brings together the wavefunction and the operators and

    establishes the link between formal calculations and experimental observations:

    Postulate 3. When a system is described by a wavefunction c, the meanvalue of the observable O in a series of measurements is equal to the expec-tation value of the corresponding operator.

    The expectation value of an operator O for an arbitrary state c is denoted hOiand defined as

    hOi ¼Rc�Oc dtRc�cdt

    ¼ hcjOjcihcjci ð1:23Þ

    If the wavefunction is chosen to be normalized to 1, then the expectation

    value is simply

    hOi ¼Z

    c�Oc dt ¼ hcjOjci ð1:24Þ

    Unless we state otherwise, from now on we shall assume that the wave-

    function is normalized to 1.

    . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

    2. This prescription excludes intrinsic observables, such as spin (Section 4.8).

    20 j 1 THE FOUNDATIONS OF QUANTUM MECHANICS

  • The meaning of Postulate 3 can be unravelled as follows. First, suppose

    that c is an eigenfunction of O with eigenvalue o; then

    hOi ¼Z

    c�Oc dt ¼Z

    c�oc dt ¼ oZ

    c�c dt ¼ o ð1:25Þ

    That is, a series of experiments on identical systems to determine O will givethe average value o (a real quantity, because O is hermitian). Now supposethat although the system is in an eigenstate of the hamiltonian it is not in an

    eigenstate of O. In this case the wavefunction can be expressed as a linearcombination of eigenfunctions of O:

    c ¼X

    n

    cncn where Ocn ¼ oncn

    In this case, the expectation value is

    hOi ¼Z X

    m

    cmcm

    !�OX

    n

    cncn

    !dt ¼

    Xm, n

    c�mcn

    Zc�mOcn dt

    ¼Xm,n

    c�mcnon

    Zc�mcn dt

    Because the eigenfunctions form an orthonormal set, the integral in the last

    expression is zero if n 6¼m, is 1 if n¼m, and the double sum reduces to asingle sum:

    hOi ¼X

    n

    c�ncnon

    Zc�ncn dt ¼

    Xn

    c�ncnon ¼X

    n

    cnj j2on ð1:26Þ

    That is, the expectation value is a weighted sum of the eigenvalues of O,the contribution of a particular eigenvalue to the sum being determined by the

    square modulus of the corresponding coefficient in the expansion of the

    wavefunction.

    We can now interpret the difference between eqns 1.25 and 1.26 in the

    form of a subsidiary postulate:

    Postulate 30. When c is an eigenfunction of the operator O, the determina-tion of the property O always yields one result, namely the correspondingeigenvalue o. The expectation value will simply be the eigenvalue o. When cis not an eigenfunction of O, a single measurement of the property yieldsa single outcome which is one of the eigenvalues of O, and the probability thata particular eigenvalue on is measured is equal to jcnj2, where cn is thecoefficient of the eigenfunction cn in the expansion of the wavefunction.

    One measurement can give only one result: a pointer can indicate only one

    value on a dial at any instant. A series of determinations can lead to a series of

    results with some mean value. The subsidiary postulate asserts that a mea-

    surement of the observable O always results in the pointer indicating one ofthe eigenvalues of the corresponding operator. If the function that describes

    the state of the system is an eigenfunction of O, then every pointer reading isprecisely o and the mean value is also o. If the system has been prepared in astate that is not an eigenfunction of O, then different measurements givedifferent values, but every individual measurement is one of the eigenvalues of

    1.11 THE OUTCOME OF MEASUREMENTS j 21

  • O, and the probability that a particular outcome on is obtained is determinedby the value of jcnj2. In this case, the mean value of all the observations is theweighted average of the eigenvalues. Note that in either case, the hermiticity

    of the operator guarantees that the observables will be real.

    Example 1.6 How to use Postulate 3 0.

    An operator A has eigenfunctions f1, f2, . . . , fn with corresponding eigenvalues

    a1, a2, . . . , an. The state of a system is described by a normalized wavefunction

    c given by

    c ¼ 12 f1 � 38

    �1=2

    f2 þ 38 i

    �1=2

    f3

    What will be the outcome of measuring the observable A?

    Method. First, we need to determine if c is an eigenfunction of the operator A.If it is, then we shall obtain the same eigenvalue of A in every measurement.

    If it is not, we shall obtain different values in a series of different measure-

    ments. In the latter case, if we have an expression for c in terms of theeigenfunctions of A, then we can determine what different values are possible,

    the probabilities of obtaining them, and the average value from a large series

    of measurements.

    Answer. To test whether c is an eigenfunction of the operator A we proceed asfollows:

    Ac ¼ A 12 f1 � 38

    �1=2

    f2 þ 38 i

    �1=2

    f3h i

    ¼ 12a1f1 � 38

    �1=2

    a2f2 þ 38 i

    �1=2

    a3f3 6¼ constant� c

    Therefore, c is not an eigenfunction of A. However, because c is a linearcombination of f1, f2, and f3 we will obtain, in different measurements,

    the values a1, a2, and a3 (the eigenvalues of the eigenfunctions of A that

    contribute to c). The probabilities of obtaining a1, a2, and a3 are, respectively,14,

    38, and

    38. The average value, given by eqn 1.26, is

    hAi ¼ 14 a1 þ 38 a2 þ 38 a3Comment. The normalization of c is reflected in the fact that the probabilitiessum to 1. Because the eigenfunctions f4, f5, . . . do not contribute here to c,there is zero probability of finding a4, a5, . . . .

    Self-test 1.6. Repeat the problem using c ¼ 13 f2 þ ð79Þ1=2f4� 13 if5:

    [hAi ¼ 19 a2 þ 79 a4 þ 19 a5

    1.12 The interpretation of the wavefunction

    The next postulate concerns the interpretation of the wavefunction itself, and

    is commonly called the Born interpretation:

    Postulate 4. The probability that a particle will be found in the volume

    element dt at the point r is proportional to jc(r)j2dt.

    22 j 1 THE FOUNDATIONS OF QUANTUM MECHANICS

  • As we have already remarked, in one dimension the volume element is dx.In three dimensions the volume element is dxdydz. It follows from this

    interpretation that jc(r)j2 is a probability density, in the sense that ityields a probability when multiplied by the volume dt of an infinitesimalregion. The wavefunction itself is a probability amplitude, and has no direct

    physical meaning. Note that whereas the probability density is real and non-

    negative, the wavefunction may be complex and negative. It is usually con-

    venient to use a normalized wavefunction; then the Born interpretation

    becomes an equality rather than a proportionality. The implication of the

    Born interpretation is that the wavefunction should be square-integrable;

    that isZjcj2 dt

  • In one dimension the equation has the form

    HC ¼ � �h2

    2m

    q2Cqx2þ VðxÞC ¼ i�h qC

    qt

    Equations of this form can be solved by the technique of separation of

    variables, in which a trial solution takes the form

    Cðx, tÞ ¼ cðxÞyðtÞ

    When this substitution is made, we obtain

    � �h2

    2my

    d2cdx2þ VðxÞcy ¼ i �hc dy

    dt

    Division of both sides of this equation by cy gives

    � �h2

    2m

    1

    cd2cdx2þ VðxÞ ¼ i �h 1

    ydydt

    Only the left-hand side of this equation is a function of x, so when x changes,

    only the left-hand side can change. But as the left-hand side is equal to the

    right-hand side, and the latter does not change, the left-hand side must be

    equal to a constant. Because the dimensions of the constant are those of an

    energy (the same as those of V), we shall write it E. It follows that the time-dependent equation separates into the following two differential equations:

    � �h2

    2m

    d2cdx2þ VðxÞc ¼ Ec ð1:29aÞ

    i�hdydt¼ Ey