QM PDF

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 310

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/310604584

Lectures on Quantum Mechanics

Research · November 2016


DOI: 10.13140/RG.2.2.15913.39528

CITATIONS READS
0 123

1 author:

Frank Tabakin
University of Pittsburgh
110 PUBLICATIONS   3,093 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Ant-proton Physics View project

Hyperon and Hypernuclear Physics View project

All content following this page was uploaded by Frank Tabakin on 21 November 2016.

The user has requested enhancement of the downloaded file.


QUANTUM MECHANICS
PHYSICS 265/266
LECTURE NOTES
FALL 2004–2005
Tuesday, & Thursday 9:30–10:45 a.m.
Room 319 AH.

Frank Tabakin
Department of Physics and Astronomy
University of Pittsburgh
Pittsburgh, PA, 15260
Contents

1 Historical and Experimental Origins 1


1.1 Particle Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Classical Electromagnetic Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.3 Fundamental Assumptions of Classical Physics . . . . . . . . . . . . . . . . . . . . . 3
1.4 Relativity and Classical Physics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.5 New Experiences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.6 Failure of Classical Physics, The Concept of Quanta, The Wave–Particle Paradox,
and The Measurement Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.7 Black Body Radiation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.7.1 Planck’s Thermodynamic Deduction of the Radiation Law . . . . . . . . . . . 7
1.7.2 Planck’s Statistical Mechanics Derivation . . . . . . . . . . . . . . . . . . . . 9
1.7.3 Planck’s Radiation Law - Simple 1913 Version . . . . . . . . . . . . . . . . . 10
1.8 Einstein and Black-Body Radiation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.8.1 Einstein’s Photon Gas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.8.2 Einstein’s Derivation of ρ(ν, T ) - Detailed Balance . . . . . . . . . . . . . . . 13
1.9 Specific Heat -Solids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
1.9.1 Classical Theory – DuLong and Petit . . . . . . . . . . . . . . . . . . . . . . 14
1.9.2 Einstein’s Theory of Specific Heat . . . . . . . . . . . . . . . . . . . . . . . . 15
1.9.3 Debye’s Theory of Specific Heat . . . . . . . . . . . . . . . . . . . . . . . . . 16
1.10 Photoelectric and Compton Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
1.11 Compton Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
1.12 The Atomic System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
1.13 Bohr’s Atom - Postulates (1913) – Direct quotes . . . . . . . . . . . . . . . . . . . . 20
1.13.1 Origins of Matrix Mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
1.13.2 Additional Support for Bohr’s Stationary States . . . . . . . . . . . . . . . . 22
1.14 Failures of Bohr’s Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
1.14.1 Electron Waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
1.14.2 Resolution of the Wave-Particle Paradox . . . . . . . . . . . . . . . . . . . . . 24
1.15 The Measurement Process - Uncertainty Principle . . . . . . . . . . . . . . . . . . . 24
1.15.1 The Uncertainty Principle - Heisenberg’s Microscope . . . . . . . . . . . . . . 26
1.15.2 The Uncertainty Principle - Single Slit . . . . . . . . . . . . . . . . . . . . . . 29
1.15.3 The Double Slit-Complementarity & Probability . . . . . . . . . . . . . . . . 29
1.15.4 Complementarity and Probability . . . . . . . . . . . . . . . . . . . . . . . . . 32
1.15.5 Counter Examples and Other Interpretations . . . . . . . . . . . . . . . . . . 32
1.15.6 Illustrations of the Uncertainty Relations . . . . . . . . . . . . . . . . . . . . 34
1.16 Derivation of the Schrödinger Equation – Free particles . . . . . . . . . . . . . . . . 35
1.16.1 The first derivative and why “i” in the Schrödinger Equation . . . . . . . . . 35

2 Probability and Wave Packets 37


2.1 Statistical Interpretation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
2.2 Wave Packets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
2.2.1 General Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
2.2.2 Mathematical Construction of Wave Packets . . . . . . . . . . . . . . . . . . 38
2.2.3 Some Features of Wave Packets . . . . . . . . . . . . . . . . . . . . . . . . . . 41
2.2.4 One Dimensional Wave Packets . . . . . . . . . . . . . . . . . . . . . . . . . . 41
CONTENTS ii

2.2.5 Group Velocity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41


2.2.6 Estimate of Spreading . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
2.2.7 Review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
2.3 Gaussian Wave Packet . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
2.3.1 Greens Function - Gaussian Wave Packet . . . . . . . . . . . . . . . . . . . . 47
2.3.2 Gaussian Packet: (continued) . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

3 Schrödinger Equation–Free Particle Motion 49


3.1 Probability Flux . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
3.1.1 Electron-charge . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
3.1.2 Plane Wave . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
3.1.3 Flux for a Wave Packet . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
3.2 Probability and Normalization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
3.3 Expectation Values . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
3.4 Momentum space wave function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
3.5 Expectation Values and Commutators . . . . . . . . . . . . . . . . . . . . . . . . . . 56
3.6 Commutators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
3.7 Dirac Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
3.8 Operator Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
3.9 Time Derivative Operator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
3.10 Commutators and Poisson Brackets . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
3.11 Review of Classical Mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
3.12 The Uncertainty Principle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
3.12.1 Generalization to [A, B] = i C . . . . . . . . . . . . . . . . . . . . . . . . . . 67
3.12.2 Minimum Uncertainty Wave Function . . . . . . . . . . . . . . . . . . . . . . 69
3.13 The Classical Limit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
3.14 Stationary States . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
3.14.1 Eigenfunctions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
3.15 Conditions on ψE (r): . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
3.15.1 Infinite Potentials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
3.16 Energy Eigenvalues - Bound State Eigenvalue Problems . . . . . . . . . . . . . . . . 79
3.16.1 Proof of general orthogonality of nondegenerate levels for Hermitian Hamilto-
nians . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
3.17 Stationary States - The Continuum . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
3.17.1 The Continuum - Plane Wave Basis . . . . . . . . . . . . . . . . . . . . . . . 82
3.17.2 Momentum Space - Plane Waves Stationary States . . . . . . . . . . . . . . 84
3.18 Complete Orthonormal Basis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
3.18.1 Probability Amplitudes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
3.18.2 Projective measurement and Wavefunction Collapse . . . . . . . . . . . . . . 89
3.18.3 Matrix Mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90

4 Motion in Three Dimensions 91


4.1 The Two-Body Problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
4.2 Center of Mass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
4.3 Orbital Angular Momentum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
4.4 Central Potentials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
4.5 Orbital Angular Momentum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
4.5.1 Commutation Rules for Angular Momentum . . . . . . . . . . . . . . . . . . 101
4.5.2 Dirac Notation for Spherical Harmonics . . . . . . . . . . . . . . . . . . . . . 102
4.6 Quantization of Angular Momentum. . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
4.6.1 Explicit Representation of L+ Step: . . . . . . . . . . . . . . . . . . . . . . . 105
4.7 The Radial Equation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
4.7.1 Spherical Bessel Functions - Free Radial Waves. . . . . . . . . . . . . . . . . . 106
4.8 Spherical Harmonics - Complete Set of Angle Functions. . . . . . . . . . . . . . . . . 107
CONTENTS iii

5 Bound States 110


5.1 The Hydrogen Atom. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
5.1.1 The Bound State - Radial Equation. . . . . . . . . . . . . . . . . . . . . . . . 110
5.2 Wave Function For Hydrogen. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
5.2.1 Normalization - Generating Functions . . . . . . . . . . . . . . . . . . . . . . 114
5.3 Harmonic Oscillator in One, Two and Three Dimensions . . . . . . . . . . . . . . . . 115
5.3.1 The Radial Differential Equation . . . . . . . . . . . . . . . . . . . . . . . . . 115
5.3.2 The Differential Equation and Ladder Operators for 1-D Oscillator . . . . . . 117
5.3.3 The Creation and Annihilation Operator description of a 1-D Oscillator . . . 119
5.3.4 Two and Three Dimensional Harmonic Oscillators . . . . . . . . . . . . . . . 122
5.3.5 Hydrogenic Momentum Space Wave Functions . . . . . . . . . . . . . . . . . 123
5.3.6 Glauber States . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
5.3.7 Normalization - Generating Functions . . . . . . . . . . . . . . . . . . . . . . 125

6 Scattering 127
6.1 Stationary States for Scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
6.1.1 Integral Equation For Scattering . . . . . . . . . . . . . . . . . . . . . . . . . 131
6.1.2 Formal Version of Stationary State Scattering . . . . . . . . . . . . . . . . . . 135
6.1.3 The Transition Matrix and the Lippmann-Schwinger Equation . . . . . . . . 138
6.1.4 The Optical Theorem - Stationary State Version . . . . . . . . . . . . . . . . 142

7 Scattering-Continued 146
7.1 Wave Packet Description of Scattering . . . . . . . . . . . . . . . . . . . . . . . . . . 146
7.1.1 Characteristic Lengths . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146
7.2 Scattering Wave Packets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
7.2.1 Cross Section & Wave Packets . . . . . . . . . . . . . . . . . . . . . . . . . . 153
7.3 Optical Theorem & Wave Packets . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153
7.4 Partial Wave Expansion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156
7.4.1 Transformation from Center of Mass to Laboratory Systems . . . . . . . . . . 156
7.4.2 Plane Wave Decomposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159
7.4.3 Stationary State Expanded in Partial Waves . . . . . . . . . . . . . . . . . . 160
7.4.4 Partial Wave Expansion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162
7.4.5 Solutions of the Radial Differential Equation . . . . . . . . . . . . . . . . . . 163
7.4.6 Optical Theorem–Partial Waves . . . . . . . . . . . . . . . . . . . . . . . . . 165
7.4.7 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
7.4.8 Meaning of Phase Shifts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
7.5 Resonances . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
7.6 Time Delay and Resonance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
7.6.1 Wave Functions and Resonances . . . . . . . . . . . . . . . . . . . . . . . . . 175
7.6.2 Resonances - Example of Separable Potential . . . . . . . . . . . . . . . . . . 176

8 Spin 180
8.1 Introductory Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180
8.2 Spin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180
8.2.1 Stern-Gerlach Experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181
8.2.2 Magnetic Moment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184
8.2.3 Spin of an Electron . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187
8.3 Pauli-Spin Operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
8.3.1 Pauli-Spin Matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190
8.4 Complete Orthonormal Spin-States . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190
8.4.1 Separable and Nonseparable Space-spin States . . . . . . . . . . . . . . . . . 191
8.4.2 Two Spin-1/2 Particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 194
8.4.3 Exchange Operators - Spin . . . . . . . . . . . . . . . . . . . . . . . . . . . . 196
8.4.4 Scattering With Spin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 197
8.4.5 Additional Comments about SGE . . . . . . . . . . . . . . . . . . . . . . . . 198
8.5 Basics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 200
8.6 Postulates of Quantum Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201
CONTENTS iv

8.6.1 Operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205


8.6.2 Projection operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206
8.6.3 Hermitian Conjugate or Adjoint Operator . . . . . . . . . . . . . . . . . . . . 207
8.6.4 The Algebra of Measurement - A Brief Visit . . . . . . . . . . . . . . . . . . . 212

9 Symmetries 215
9.1 Spatial Translation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215
9.2 Rotations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 218
9.2.1 Spin and Total Angular Momentum . . . . . . . . . . . . . . . . . . . . . . . 219
9.2.2 The Rotation Group . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220
9.2.3 Invariance under Rotation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220
9.2.4 Translation and Operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221
9.2.5 Rotation and Operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223
9.2.6 Angular Momentum Commutators . . . . . . . . . . . . . . . . . . . . . . . . 225
9.2.7 Addition of Angular Momentum . . . . . . . . . . . . . . . . . . . . . . . . 226
9.2.8 Addition of Three Angular momenta ( Racah Coefficients or 6-j Symbols) . . 232
9.2.9 Addition of Four Angular momenta ( 9-j Symbols) . . . . . . . . . . . . . . . 233
9.2.10 Matrix Representation of the Rotation Group . . . . . . . . . . . . . . . . . 234
9.2.11 Euler Angles and Finite Rotations . . . . . . . . . . . . . . . . . . . . . . . . 235
9.2.12 Wigner Rotation Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237
9.2.13 Spherical Harmonics and The Wigner Rotation Matrix . . . . . . . . . . . . 240
9.2.14 Clebsch-Gordan Reduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 241
9.2.15 Irreducible Spherical Tensors . . . . . . . . . . . . . . . . . . . . . . . . . . . 242
9.2.16 The Wigner-Eckart Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . 247

10 Quantum Dynamics 251


10.1 Stationary State Perturbation Theory . . . . . . . . . . . . . . . . . . . . . . . . . . 251
10.1.1 The Secular Determinant . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 252
10.1.2 Symmetry and Reduction of the Secular Equation . . . . . . . . . . . . . . . 252
10.1.3 Rayleigh-Schrödinger Perturbation Theory . . . . . . . . . . . . . . . . . . . 255
10.2 Degenerate Bound State Perturbation Theory . . . . . . . . . . . . . . . . . . . . . . 256
10.3 Application of Time-independent Perturbation Theory . . . . . . . . . . . . . . . . . 260
10.3.1 The Fine Structure of Hydrogen . . . . . . . . . . . . . . . . . . . . . . . . . 260
10.3.2 Perturbations of the Hydrogen Atom . . . . . . . . . . . . . . . . . . . . . . 262
10.3.3 Fine Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 264
10.3.4 The Zeeman and Paschen-Back Effects . . . . . . . . . . . . . . . . . . . . . 265
10.4 Time Dependent State Perturbation Theory . . . . . . . . . . . . . . . . . . . . . . . 271
10.5 Time Evolution in Quantum Mechanics . . . . . . . . . . . . . . . . . . . . . . . . . 271
10.5.1 Schrödinger and Heisenberg Pictures . . . . . . . . . . . . . . . . . . . . . . . 273
10.5.2 Dirac or Interaction Picture . . . . . . . . . . . . . . . . . . . . . . . . . . . . 275
10.5.3 The Dyson Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 276
10.5.4 Perturbation theory–“Weak” V (t). . . . . . . . . . . . . . . . . . . . . . . . . 278
10.6 The Free Electromagnetic Field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 283
10.6.1 Quantization of the Electromagnetic Field . . . . . . . . . . . . . . . . . . . . 286
10.6.2 The Uncertainty Principle For Fields . . . . . . . . . . . . . . . . . . . . . . . 287
10.6.3 The Coupled System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 291
10.6.4 Radiation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 292
10.7 The Dirac Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 298
10.7.1 References on the Dirac Equation . . . . . . . . . . . . . . . . . . . . . . . . . 298
10.7.2 The Dirac Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 299
10.8 Derivation of the Dirac Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 301
CONTENTS v

PHYSICS 265 – FALL, 2004


LECTURE SCHEDULE
29 LECTURES IN 14 WEEKS

NOS. DATE TOPIC


1. Aug. 31 T Review of Historical and Experimental Basis of Quantum Mechanics
2. Sept. 2 TH Planck, Black–Body Radiation and Quanta
3. Sept. 7 T Einstein, Black–Body Radiation, Photoelectric, Compton, & Photons
4. Sept. 9 TH Bohr, Atoms, Correspondence Principle, Rydberg–Ritz, Heisenberg
5. Sept. 14 T Wave–Particle Duality & Complementarity
6. Sept. 16 TH The Measurement Process–Uncertainty Principle
7. Sept. 21 T Wave Packets–Schrödinger Equation;Probability Flux–Expectation Values
8. Sept. 23 TH Operators–Dirac Notation
9. Sept. 28 T Hermitian, Adjoint Operators;Commutators–Time derivative Operator
10. Sept. 30 TH Operator Equation of motion–The Classical Limit
11. Oct. 5 T Review of Classical Physics–Poisson Brackets
12. Oct. 7 TH The Uncertainty Principle and Commutators
13. Oct. 12 T Eigenfunctions, Stationary States
14. Oct. 14 TH Boundary Conditions;Bound and Continuum States - Plane Waves - CON Sets
15. Oct. 19 T Box Normalization, Complete Orthonormal Basis - Closure - Dirac Notation
16. Oct. 21 TH Probability Amplitudes, Matrix Mechanics - Two Body Problem
17. Oct. 26 T Radial Schrödinger Eq’n - Angular Momentum
18. Oct. 28 TH Angular Momentum - Spherical Harmonics - Free Particles in 3D-Bessel Functions
19. Nov. 2 T Coulomb Problem - Bound State Problems- Laguerre Polynomials
20. Nov. 4 TH Scattering - Stationary State Theory
21. Nov. 9 T Stationary State Theory - Greens Functions - Born Approx
22. Nov. 11 TH Lippmann Schwinger Equation
23. Nov. 16 T Time-Dependent - Wave Packet Theory of Scattering– Characteristic Lengths
24. Nov. 18 TH Cross Section and Optical Theorem
25. Nov. 23 T Plane waves;Partial Wave Expansion - Radial Integral Equation
26. Nov. 30 TH Partial Wave Lippmann Schwinger Equation- Phase Shifts, Cross Sections
27. Dec. 2 T T-Matrix - Resonance - Solutions
28. Dec. 7 T Analytic Properties - Bound States - Resonances - Thresholds
29. Dec. 9 TH Review - Summary - Next Term Topics

Tentative Exam Dates:


Exam # 1 Sat. October 16th.
Exam # 2 Sat. November 20th.
FINAL EXAM During week of Dec. 13th, 2004.
CONTENTS vi

BIBLIOGRAPHYa

History
1. William H. Cropper “The Quantum Physicists”, Oxford University Press.
2. B. L. Van der Waerden “Sources of Quantum Mechanics”, Dover Publications,Inc.

3. Max Jammer “The Conceptual Development of Quantum Mechanics ” McGraw–Hill.


4. A. P. French and P. J. Kennedy “Niels Bohr – A Centenary Volume”, Harvard University
Press.
5. A. Hermann “ The Genesis of Quantum Theory (1899 – 1913)”, M.I.T. Press.

6. “Niels Bohr and the Development of Physics”, Edited by W. Pauli, Pergamon Press.
7. Niels Bohr “Atomic Physics and Human Knowledge”, Sc. Ed. – 1961.
8. “Albert Einstein - Philosopher–Scientist”, Edited by P. Schilpp.
9. P. Stehle “Order, Chaos, Order ”,Oxford University Press –1994

Texts
1. A. Messiah “Quantum Mechanics” The $ 29 Dover edition is available in the bookstore
2. J. J. Sakurai “Modern Quantum Mechanics”, Addison–Wesley Publishing Co.

3. L. I. Schiff “Quantum Mechanics”


4. P. A. M. Dirac “The Principles of Quantum Mechanics”
5. Gottfried, Kurt and Yan, Tung-Mow “Quantum Mechanics: Fundamentals,” 2nd ed., 2004.

aI will be adding references to this as the term proceeds.


Chapter 1

Historical and Experimental


Origins

Let me begin our study of quantum mechanics by reviewing the historical and experimental origins.
I emphasize that it is a review since I suspect much of what I will say has already been seen by you.
Nevertheless, I believe it is important to firmly understand these origins in order to fully appreciate
the scope of quantum theory.
At the beginning of the 20th century, theoretical physics seemed to be almost complete. The
two basic subjects of particle dynamics (Newtonian Mechanics) and electromagnetic field theory
(Maxwell) were nearly perfect.

1.1 Particle Motion


The basic notion of classical physics was that the motion of bodies can be described in terms of
orbits or trajectories. A particle view was adopted for bodies that could be localized in space (or
thought of that way by using the center of mass). The location and motion of a particle is then
described by defining its space (~r) and time (t) coordinates and its velocity (~v ). Given the initial
location ~r(t0 ) and initial velocity ~v (t0 ) at time t0 , one could predict with arbitrarily good precision
the subsequent development of the trajectory; i.e.,
Z t ~ 0
F (t ) 0
~v (t) = ~v (t0 ) + dt ,
t0 m
given the applied force. Also
t00
t
F~ (t0 ) 0
Z Z
~r(t) = ~r(t0 ) + ~v (t0 )(t − t0 ) + dt00 dt .
t0 t0 m
There are several concepts used here that I wish to mention, because they will be called into
question by quantum theory.
1. The notion of a particle is based on a body being considered localized in space, without its
shape, color, temperature, etc. influencing its motion - its motion is affected only by its mass
and the applied force.

t
B

t0
A

Figure 1.1: A Trajectory


CHAPTER 1. HISTORICAL AND EXPERIMENTAL ORIGINS 2

2. The measurement of the particle’s location and velocity (as well as the acceleration and applied
force) could be carried out with any precision desired. The object’s properties are determined
and well separated from the properties of the measuring device. A clean, precise separation of
the objective and subjective is basic to classical physics. The measuring device can be made
to have a negligible effect on the system itself, or at least the effect of the measurement can
be corrected for using known laws of dynamics.
3. With each set of physical attributes as determined by measurements, there are a set of numbers
(meter or speedometer readings, etc) that are used to define the state of the system.

1.2 Classical Electromagnetic Fields


In addition to particle dynamics, the subject of electromagnetic fields forms the second great classical
concept; namely, the Faraday concept of fields. A field is understood to be spread out or distributed
~ r, t). The
in space. For example, the electric field is described as a function of space and time as E(~
electric and magnetic fields propagate according to a wave equation (in the absence of currents) as
a
1 ∂2 ~
2A(~
~ r, t) = 0 = (∇2 − )A(~r, t) = 0. (1.1)
c2 ∂t2
The idea of a field is that it is extended (or spread-out) in space, in contrast to a particle which can
be thought of as localized in space. The most “spread out” case is simply the plane wave solution
of Equation 1.1, namely
~ = Â(~k) ei(~k·~r−ωt)
A (1.2)

with ~k = and ω = 2πf , and, by virtue of c = f λ, we have
λ
ω = kc. (1.3)

By c we mean the “phase velocity” of the above plane wave b , e.g. vphase = ω/k.
We can, however, consider fields that are not so spread out as in Equation 1.2. Instead, we can
form a pulse, or “wave packet” which is localized to some extent; i.e.
~
ei(k· ~r−ω t) ~
Z
~ t) =
A(r, Â(~k − ~k0 ) dk
(2π)3/2

is also a solution of Equation 1.1 since ω = kc, but it can be made localized by appropriately
mixing, or superposition of waves with various wave lengths. The wave packet travels with the
~ k ω(k) |k (or for the 1-D case: vgroup = ∂ω |k .) I shall have more to
“group velocity” ~vgroup = ∇ 0
∂k 0
say about such wave packets later. For now, I only wish to emphasize that waves, or fields, are
spread out in space, although they can be localized to some limited extent by forming such “wave
packets,” or pulses.
As a consequence of fields being “spread out” in space, they can produce the interference and
diffraction patterns characteristic of wave motion. It is the existence of such patterns that forms
the most convincing evidence that light is a wavec .
It is also worth noting that the electromagnetic field carries energy, ∼ ( E 2 + B 2 ), and
momentum ∼ (E ~ × B),
~ which are concepts that originate from particle dynamics. In this regard,
and also when considering wave packets of “small” size (i.e., small compared to our experimental
resolution), there appears to be some cases where fields can be approximately described in terms of
a Recall ~ r, t) = − 1 ∂ A
that: E(~ ~ − ∇Φ
~ and B(~
~ r, t) = ∇
~ × A;
~ often we take Φ = 0.
c ∂t
b Forlight in vacuum ω = kc and then the phase velocity and group velocity both equal “c”. For a dispersive
medium, ω 6= kc and the two velocities differ. That dispersive case also holds for the Schrödinger equation, as we will
see later.
c The early particle theories of light (Newton, Laplace and Biot) were overthrown by Fizeau and Foucault’s deter-

mination of the speed of light being smaller in a medium, rather than larger as the particle theory required to explain
refraction. The wave theory was established by the phenomena of interference and diffraction (Huyghens, Young and
Fresnel); the first recorded observations of diffraction were by da Vinci(1452–1519), and Grimaldi(1665).
CHAPTER 1. HISTORICAL AND EXPERIMENTAL ORIGINS 3

“particle–like” motion. In fact, we know that ray optics or geometrical optics (so familiar from lens
studies) does make use of the notion of a trajectory or rays in analogy to particle dynamicsd .
That fact so intrigued W. R. Hamilton (Scottish; 1805-1865) that he reformulated mechanics
in terms of a variational principle and a “Hamiltonian,” to emphasize that mechanics could be
thought of as an analogy to geometrical optics (and vice-versa). Despite this heroic analogy, the
fact remained that in classical physics a field is a field (i.e., it displays interference and diffraction
in the proper circumstances λ ≈ a) and a particle is a particle (it always has a precise trajectory,
albeit undetermined for many-body systems). For many-particles systems one turns to statistical
mechanics which nevertheless assumes the underlying existence of definite trajectories.

1.3 Fundamental Assumptions of Classical Physics


Over and above the dynamical laws (F = ma and Maxwell’s equations) which determine the behavior
of particles and fields, there are several fundamental assumptions made in classical physics that are
called into question by quantum theory. I have already mentioned some of these assumptions, let
me provide a list:

1. One can measure the location (~r), speed (~v ), and acceleration (~a) of particles and the field
~ B,
properties (E, ~ E ~ × B,
~ E 2 + B 2 ) with sufficiently high precision so as to determine the
objective reality of the system with the role of the (subjective) observer made negligible. That
is: it is assumed conceptually possible to devise an apparatus which will measure without
inducing large perturbing effects on the object under study. In addition, any effects on the
system due to the measurement process can be corrected for using known dynamical principles
and thus the unperturbed nature of the system can be extracted.
2. There is always the possibility of making infinitesimally small changes in the attributes of a
classical system. (For example, ∆~r, ∆t, ∆v, ∆m, ∆ee or ∆E, ∆B etc.) Even in cases where
discreteness occurs (a string of length L with fixed ends) it is always possible to change L and
hence the harmonics by infinitesimal amounts. Therefore, continuity is a basic characteristic
of classical theories and in fact, is essential for the design of “unperturbing” or ”correctable”
measurements as discussed in 1) above.

1.4 Relativity and Classical Physics


It is the effort to understand electromagnetic waves (and their interaction with particles) that caused
the above classical concepts to come under intense scrutiny. It should be noted, however, that in a
fundamental way, the theory of relativity is still a “classical theory.” Einstein did call into question
the validity of earlier conceptual ways of measuring space and time. The natural limitation of our
ability to determine simultaneity is provided by the large, but finite speed of light c = 3 × 1010
cm/sec.
Einstein showed that new realms of experience (higher speeds) requires reconsideration of “time-
honored” concepts – he invoked the procedures of “operational definition” described in a unique
reasoning method called “gedanken experimenten” or thought experiments. These methods beau-
tifully illustrated the conceptual limitations of Newtonian mechanics. (He also taught us to avoid
concepts that are not accessible to measurements, such as the aether.)
However, despite these profound changes in our concepts and the new methods of thought pro-
vided by relativity (much of which were adopted by Bohr and Heisenberg in developing quantum
mechanics), there is a fundamentally “classical” notion used in relativity. Namely, the observation
of ~r, ~v , ~a could be carried out without disturbing the system –trajectories existed and in fact, the
objective and subjective worlds exist and can be separated out.
Therefore, both 1) and 2) are concepts that are unchanged by introduction of the limiting speed
c into physics. Thus, the quantum theory, which questions 1) and 2), provides even a greater
d Geometrical or ray optics is the short wave limit of wave optics. For a slit (or wave packet) of size a, we have

ray optics for the limit λ  a ; introducing, the DeBroglie wave length–momentum relation, λ = h p
, the above limit
becomes pa  h . We already see that the classical limit of ”rays” obtains for large pa or for h → 0.
e The existence of discrete masses and quantized charges for elementary particles is already a violation of classical

ideas.
CHAPTER 1. HISTORICAL AND EXPERIMENTAL ORIGINS 4

revolution in our concepts than is provided by relativity. Now I would like to present the reasons
for questioning concepts 1) and 2).

1.5 New Experiences


We learn from Einstein that new experiences require reconsideration of time–honored concepts. The
“new experiences” of 1900 – 1925 showed that classical physics failed. Furthermore, the “wave-
particle duality” was demonstrated. The “wave particle duality” (also called complementarity, see
later) refers to the observation that
• a “particle” (electron) sometimes displays wave properties (interference and diffraction)
and
• a “wave” (light) sometimes displays particle properties (a photon-transfers energy and mo-
mentum as a particle)
The evidence for these facts arose in new realms, the atomic realm, which are very small ∼ 10−8
cm, compared to the realms of previous experience. The notion of “small” will be made more explicit
later (in terms of h̄/mc), for now let me just point out “small” is not determined by extent in space
(h̄/mc) only, but will be determined by the “action” involved in the process. Action is ∼ energy
× time interval, f and has the units ML2 /T - as does Planck’s constant h̄. Therefore, h̄, which
is a universal constant of nature, will serve to measure the “small-scale” nature of a system. For
example, we will see that for classical systems the action involved in the motion is very much greater
than h̄ and one can neglect the finite size of h̄ (The classical limit is then taken by either setting
h̄ → 0 or going to cases with large action.) Quantum mechanics treats “small-scale” systems like
atoms, for which the action is ∼ h̄.
Next we shall see how “new experiences” caused the failure of classical physics and the alteration
of its fundamental assumptions.

f Action is also momentum× coordinate or p · dq


CHAPTER 1. HISTORICAL AND EXPERIMENTAL ORIGINS 5

1.6 Failure of Classical Physics, The Concept of Quanta, The


Wave–Particle Paradox, and The Measurement Process
Reference: W. H. Cropper The Quantum Physicists and an Introduction to their Physics–on reserve

The above topics will now be discussed. The failure of classical physics is seen clearly in studying:
(1) black-body radiation,(2) in the photoelectric effect, (3) in the stability of atoms, etc., which
simultaneously serve to establish the existence of photons, or quanta of energy. These discrete
aspects also occur in Bohr’s model.
Having established the existence of quanta in nature and the corresponding breakdown of classical
concepts, we will develop the evidence for the dual wave–particle nature of both light and matter
(electrons). That paradox of having both wave and particle properties, will then be resolved by
studying the measurement process and the restrictions imposed by nature on our ability to measure
- i.e., as stated in the uncertainty principle. Consideration of the double slit will lead us to the
probabilistic interpretation of quantum mechanics and on to the probability amplitude waves. The
topics are to be discussed in the order:

• Black–Body Radiation
a) existence of h̄ - Planck
b) Einstein demonstration of photons

• Photoelectric and Compton Effects


• The Atomic System
a) instability of classical theory
b) Bohr’s model - postulates and correspondence principle
c) limitations of Bohr’s theory
d) Ritz combination principle
• Discreteness in Space: Stern–Gerlach
• Electron Waves–DeBroglie

• Wave Particle Paradox


• The Measurement Process
a) Uncertainty Principle
b) Probability waves
c) Complementary

After these topics are discussed, we’ll turn to wave packets, wave functions, Schrödinger’s equa-
tion, operators, mathematical proofs of the uncertainty principles, matrix elements and the Heisen-
berg or matrix mechanics approach–followed by the general transformation theory formulation of
quantum mechanics(Dirac).
As a historical guide, let me present the following chronological chart
CHAPTER 1. HISTORICAL AND EXPERIMENTAL ORIGINS 6

YEAR DEVELOPMENT DISTANCES PEOPLE


≤ 1900 Classical Physics ≥ 10−4 cm Newton & Maxwell
1900 Old Quantum Theory and Idea of Quanta ≥ 10−8 cm Planck, Einstein, Bohr
1913 Canonical Variables–Quantization and Action ≥ 10−8 cm Bohr, Sommerfeld
1926 Quantum Mechanics ≤ 10−4 cm Bohr, Born, Heisenberg, Jordan
Dirac
1926–1935
Matrix Mechanics: Bohr, Born, Heisenberg, Jordan
Wave Mechanics: DeBroglie–Schrödinger, Bohr
Transformation Theory: Dirac, Jordan
Relativistic Quantum Mechanics: Dirac
1940– Nuclear Reactors & The Bomb = 10−13 cm Fermi, Bethe
1947 High Energy Accelerators ≤ 10−13 cm ···
1947–2003 Atomic Beams, Lamb Shift & Quantum Electrodynamics Dirac , Schwinger, Feynman
Mesons , Nucleons ,Electrons, Antiparticles
Lasers, semiconductors,superconductors,
transistors, integrated circuits
S–matrix, field theory, current algebra, group theory
symmetries,dispersion relations, Isospin,SU3,
Strangeness, Hyperons, Baryons, Hadrons,
Weak and strong unification, QCD, GUTS · · ·

Table 1.1: Chronology


CHAPTER 1. HISTORICAL AND EXPERIMENTAL ORIGINS 7

1.7 Black Body Radiation


It was Planck’s effort in 1900 to explain the properties of radiation in equilibrium with matter
that led to the introduction of Planck’s constant h and the notion of fundamental discreteness, or
quanta, into Physics. Earlier, G. Kirchhoff (1859) had proven, on general thermodynamic grounds,
that the energy spectrum of the radiation in a cavity (black–body radiation) is independent of the
shape, size, and chemical composition of the oven. (Black–body radiation corresponds to radiation
in equilibrium with a perfect absorber and emitter). Using that general result, Planck considered
the walls to be made of simple harmonic oscillators in equilibrium with their emitted radiation – by
virtue of Kirchhoff’s theorem he knew that the spectral distribution deduced this way would be a
general property of nature.
Planck’s original deduction of the spectral density of black–body radiation g
8πhν 3 hν/kT
ρ(ν, T ) = (e − 1)−1 . (1.4)
c3
was made by a brilliant interpolation using the thermodynamic law–of entropy. (That derivation
follows.) Thus, ρ(ν, T ) was established firmly on general and empirical grounds first, then a mi-
croscopic deduction was invoked which introduced quanta. The product ρ(ν, T )dν has the units of
energy/volume and correspondingly ρ(ν, T ) is in (energy/volume)×sec
Before using a statistical mechanics approach, Planck used the concept of entropy, plus some
“new” (1900) experimental results, to deduce that the radiation law is of the above form. It was
only after verifying that Equation 1.4 was the true universal radiation law for black–body radiation
that Planck turned to a microscopic theory to explain it.
Time permitting, let me present simplified versions of Planck’s derivations.

1.7.1 Planck’s Thermodynamic Deduction of the Radiation Law


References: W. H. Cropper “The Quantum Physicists” p. 11.
M. Jammer p. 18 -
Derivation
Consider the entropy of the cavity radiation in equilibrium with the wall oscillators; the radiation
and the wall oscillators have the common frequency ν and the same absolute temperature T. We
have in general for fixed volume V
dQ dE P dV
dS = = + (1.5)
T T T
dE
→ . (1.6)
T
Here E = E(ν, T ) is the average energy of the wall oscillators and T is the absolute temperature of
the walls and the cavity radiation. Now consider the second derivative of the entropy, S ,
d2 S 1 dT
= − (1.7)
dE 2 T 2 dE
dE −1
= −[T 2 ] (1.8)
dT
It was known that at high frequencies Wien’s law holds true, i.e.

ρ(ν, T ) = ρν = α ν 3 Φ(ν/T ) = α ν 3 e−β ν/T ,


where Φ(ν/T ) was originally an unknown, but universal, function of the ratio ν/T. Here α is an
empirical constant. The last part above includes a determination of Φ for the high frequency region
made by Wien using a classical Doppler frequency shift argument.
With Planck’s result that the average energy of the wall oscillators is
c3
E(ν, T ) = ρν (1.9)
8πν 2
g Here ρ(ν, T ) is the electromagnetic energy density at the temperature T ( in Kelvin degrees) within the frequency

range ν to ν + dν. Note that ρ(ν, T )dν = ρ̄(λ, T ) dλ.


CHAPTER 1. HISTORICAL AND EXPERIMENTAL ORIGINS 8

(see later), Wien’s law gives the internal energy as

ανc3 −βν/T
E= e . (1.10)

We therefore obtain from Wien’s law that at high frequency, ν:
dE βν
= 2E
dT T
or that

d2 S
00 1
S ≡ =− (1.11)
dE 2 CE
where the constant C ≡ βν is independent of T (depends on ν).
In 1900, Ruben and Kurlbaum showed experimentally that ρν (T ) ∼ T for low frequencies, ν,
and high T. Using Equation 1.9, we see that the energy E must also be proportional to T and we
have in that region the behavior
00 d2 S 1
S ≡ = − 0 2, (1.12)
dE 2 CE
where C 0 is another constant independent of the temperature, T. How can the two cases , of high
frequency and low T (where Wien’s law applies), and the case of low frequency and high T (where
the Rubens and Kurlbaum data rules), be reconciled?
Planck’s great intuitive ideah was to combine these two limiting cases using the following inter-
polating rule

[S 00 ]−1 = −E(a + E) or S 00 = −[E(a + E)]−1 (1.13)


where a depends on ν but not on T! Integration now yields first:
d dS 1
=− (1.14)
dE dE E(a + E)

and then
Z
dS dE
= − ; (1.15)
dE [E(a + E)]
Z
1 dE dE
= − ( − ), (1.16)
a E a+E
dS 1 a+E
= [ln( ) + b] (1.17)
dE a E
1
= (1.18)
T
For T → ∞, we have ρ ∝ T and E ∝ T . In that limit, the above equation reduces to
1
ln 1 + b = 0
a
and we find that the integration constant vanishes, b = 0, using the “new” (1900) results. Solving
for the internal energy, we get
a
E = a/T
e −1
and finally, the form

ρ(ν, T ) = ( 3 )ν 2 a (ea/T − 1)−1 (1.19)
c

ν
7−→ αν 3 e−βν/T ,
T →∞

h “Never in the history of physics was there such an inconspicuous mathematical interpolation with such far–reaching

physical and philosophical consequences”, see M. Jammer’s book p. 18


CHAPTER 1. HISTORICAL AND EXPERIMENTAL ORIGINS 9

where the high frequency, low temperature limit recovers Wien’s law with a = βν. The other limiting
case, of low frequency and high temperature, gives

8π βν 3 8π
ρ(ν, T ) = 3 βν/T
7− → 3 ν 2 T, (1.20)
c e − 1 T →0 c
ν

which we see reduces to the Ruben and Kurlbaum result.


An empirically solid interpolation was made by Planck, who then went on to seek its fundamental
significance.

1.7.2 Planck’s Statistical Mechanics Derivation


Here Planck’s use of discrete wall oscillator energies to derive his radiation formula is sketched.
Consider the entropy SN of N wall oscillators of frequency ν:

SN = k ln W
Here W denotes the number of distributions of basic energy “elements” () compatible with the
total energy of the system EN
EN = N E(ν, T ) = n,
where n is the number of energy elements, each of energy , to be distributed to the N wall oscillators.
Here E(ν, T ) denotes the average energy of a wall oscillator.i
To determine the entropy SN , we need W, the number of ways of distributing n energy elements
among N oscillators. Energy conservation requires that EN be fixed, which we satisfy by keeping 
and n constant. Counting the number of ways to distribute the n bundles of energy, , over the N
oscillators yields for W
(N + n − 1)! (N + n)N +n
W = ≈ , (1.21)
(N − 1)!n! N N nN
and therefore
SN = k[(N + n) ln(N + n) − N ln N − n ln n]. (1.22)
Here we used Sterling’s approximation: ln x! = x ln x − x for large x. Now using energy conservation,
from above, n/N = E/, and hence

SN E E 1+ E
S= = k[ln(1 + ) + ln E  ]. (1.23)
N   
Taking the derivative of this entropy and using its relation to the temperature of the wall ( which,
for the equilibrium we assume, equals the temperature of the radiation), we find

∂S 1
= (1.24)
∂E T
k 1+ E E 1 1
= {(1 + E/)−1 + ln E  + [ E
− E]} (1.25)
   1 +  
k 1 + E
= {ln E ] }. (1.26)
 
Therefore,

e− kT
E=  . (1.27)
1 − e− kT
With  set equal to h ν, this yields Planck’s major result that the average oscillator energy is

E= hν . (1.28)
e kT −1
i This finite counting scheme was introduced originally by Boltzmann as a way of treating continuous variables by

combinatorial methods. See Cropper [p. 14 ] for Planck’s revision (by adding in h) of Boltzmann’s derivations.
CHAPTER 1. HISTORICAL AND EXPERIMENTAL ORIGINS 10

1.7.3 Planck’s Radiation Law - Simple 1913 Version


j
~ is the vector
Electromagnetic standing-wave modes in a box satisfy the wave equation, where A
potential
2A ~ = 0, (1.29)
with appropriate boundary conditions. These modes are of the form
X
~∝
A · · · sin kx x sin ky y sin kz z,

with

ki L = ni π for ni = 0, 1, 2, · · · ≥ 0 & ki ≥ 0 (1.30)


V 3
d k = d3 n for n ≥ 0. (1.31)
π3
If we extend the range of ni and ki , to negative numbers we have a sphere in k-space and need an
extra factor of 1/8 [ this is Jean’s contribution to the Rayleigh–Jean’s law - see Jammer p. 16 !!].
Now also include the two transverse, i.e. two polarization states of A~ and the relation ω = kc =
2πν to obtain:

d3 n 1 1 1 8π
= 2 3 4πk 2 dk = 2 3 ω 2 dω = 3 ν 2 dν; (1.32)
V 8π π c c

where the number of modes in dν = 3 ν 2 dν and
vol c
energy energy mode 8πν 2
ρ(ν, T )dν = in dν = = E(ν, T ) dν ,
vol mode vol c3
which is a result we used earlier in Eq. 1.8; namely that

8πν 2
ρ(ν, T ) = E(ν, T )
.
c3
The energy E(ν, T ) is the average energy of radiation is equated to the average energy of the wall
oscillators for which the classical result is:
1
E(ν, T ) = 2 × kT = kT,
2
2
since H = p2 + q 2 has two degrees of freedom. In that case , we get ρ(ν, T ) = 8π νc3 kT, the Rayleigh-
Jeans law. Note that the Stefan-Boltzmann law for the (energy/(sec-area)) emitted by a blackbody

c ∞
Z
ρ(ν, T )dν = σT 4
4 0
diverges if the Rayleigh-Jeans law is used in the above Stefan-Boltzmann law; we have a high
frequency or so called ultra-violet catastrophe.
Turning to Wien’s law

ρ(ν, T ) ∝ ν 3 e−βν/T and E ∝ νe−βν/T

the above catastrophe is cured, but Wien’s law is known to fail for high T and low ν.k
The day is saved by Planck’s law , which was first obtained by the entropy interpolation method
and then the thermodynamic reasoning presented earlier. We have:

E(ν, T ) = hν 7−→ kT
e kT −1 h→0

j Original Methods are described in Jammer and Cropper.


k Wien also deduced his famous displacement law λmax T = b and also the general form mentioned earlier. Note
the T 4 dependence follows from the Wien’s general form ρ(ν, T ) = ρν = α ν 3 Φ(ν/T ).
CHAPTER 1. HISTORICAL AND EXPERIMENTAL ORIGINS 11

which is the classical limit, or the low frequency, high temperature limit

7−→ kT
ν
kT →0

which is the experimental result of 1900. The other limit returns the Wien result

7−→ hνe−hν/kT .
ν→large

The above result, which resolves the ultraviolet catastrophe and agrees with Wien’s law, now
requires only the derivation of Planck’s Law using the idea of quantized wall oscillators. The wall
resonators, or “atoms,” are viewed as harmonic oscillators which transfer energy in bundles, or
quanta 0, hν, 2hν · · ·. Their average energy is determined by the thermal averaging rule:

hνe−hν/kT + 2hνe−2hν/kT + · · ·
E =
1 + e−hν/kT + e−2hν/kT + · · ·
P∞
0Pn h ν e−nhν/kT
= ∞ −nhν/kT
e
P0
hν Xn
= P nn
nX
∂ X
= hνX ln Xn
∂X n
∂ 1
= hν X ln
∂X 1 − X
hνX hν
= = 1
1−X X −1

Here X ≡ e−hν/kT . This gets us again to Planck’s result for the average energy of the wall oscillators

E(ν, T ) =
ehν/kT −1
and thus
8πν 2 hν
ρ(ν, T ) = hν Planck’s Law
c3 e kT −1
Planck stressed that the wall oscillators were quantized and in thermal equilibrium with the
radiation, with the radiation still described classically. Later, A. Einstein showed that the radiation
itself behaves as a “gas” of particles later called photons. In 1916, A. Einstein used detailed balanced
for radiation and matter in equilibrium to deduce ρν ! His idea of using a quantum description of the
radiation, not just of the wall oscillators, is presented next.

1.8 Einstein and Black-Body Radiation


Einstein found Planck’s derivation unsatisfactory, because of the mixed approach used by Planck,
although Einstein believed Planck’s result to be correct. Roughly, Einstein’s objection was that the
formula

8πν 2
ρ(ν, T ) = E(ν, T )
c3
has E(ν, T ) based on discrete wall oscillator energies, while continuous classical electromagnetic
2
modes were used to derive the 8πνc3 factor , i.e. the Planck derivation has two opposite assumptions.
Einstein felt that the continuity ideas used to derive the density of modes was inconsistent.
Therefore, he deduced Planck’s formula on more general thermodynamic grounds, based on the idea
of quanta. Reference: Annalen der Physik 17 132 (1905).
Furthermore, Einstein showed that under volume changes the radiation behaved as a “gas” of
quanta or photons. He therefore attributed the discreteness to the radiation as well as to the wall
oscillators - a step that Planck never made and had strong objections to!
CHAPTER 1. HISTORICAL AND EXPERIMENTAL ORIGINS 12

Einstein showed that the entropy of the cavity radiation changed according to
R V
ln( )N Eν /βνR ,
S − S0 =
N V0
under a change of the cavity volume V0 to V . Here Eν = V ρν dν = total energy of the radiation of
frequency ν in volume V. He compared this to the kinetic theory of gas result
R V
S − S0 = ln( )n ,
N V0
where n = # particles ≡ N Eν /(βνR) and thus he identified

Eν = n( )ν = nhν
N
R
using β = h/k and N = k = the Boltzmann constant l . This wonderful result, often stated as

Eγ = h̄ω,

with h̄ = h/(2π) and ω = 2πν, relates the energy of a photon (particle–like aspect) to its frequency
(wave–like aspect). It first appeared in Einstein’s photoelectric paper of the “golden year” of 1905.

1.8.1 Einstein’s Photon Gas


The idea of cavity radiation as a photon gas was contained in the famous photoelectric paper
(Annalen der Physik 17 132 (1905), also see Jammer p. 29 ff). Consider the entropy per volume of
the cavity radiation of frequency ν
Sν /V = sν (ρν ).
Thus sν (ρν ) is the “entropy density” of radiation of frequency ν and the total entropy is given by
Z ∞ Z ∞
S= dν Sν = V dν sν (ρν ).
0 0

The total energy E is given in terms of the energy density ρν ≡ ρ(ν, T ) by


Z ∞
E=V ρ(ν, T ) dν.
0

Recall that in general


1
dE,dS =
T
which for a change in ρν by δρν , can be expressed as
Z ∞
∂sν
dS = V dν δρν .
0 ∂ρν
The incremental change in energy can also be expressed in terms of a change in ρν by
Z ∞
dE = V δρν dν.
0

Invoking a variational principle for the entropy (maximized)


Z
δ sν (ρν ) dν = 0,

and for the energy (energy conserved)


Z ∞
δ ρ(ν, T ) dν = 0,
0
l For the ideal gas, we have S = k ln V n with n = # molecules. This follows from dQ = T dS = dU + P dV and
∂S
T ∂V = P = TVkn then P V = nkT = n NR T = n0 RT . The number of moles is given by n0 .
0
CHAPTER 1. HISTORICAL AND EXPERIMENTAL ORIGINS 13

which are combined into the statement that


Z ∞
∂sν
( − λ) δρν dν = 0,
0 ∂ρν

for any δρν around any ν. Thus, the Lagrange multiplier,λ, is determined by
∂sν
λ= = independent of ν.
∂ρν
∂sν
Using the above frequency independence of , we can factor that functional derivative outside of
∂ρν
the integration over ν in the expression
Z ∞
∂sν dE
dS = V δρν dν = .
∂ρν 0 T
Hence, we conclude that
∂sν (ρν ) 1
= .
∂ρν T
From Wien’s law (good for low temperature and high ν), we have

ρ = αν 3 e−β ν/T ;

therefore,
1 1 ρ
=− ln .
T β ν αν 3
We arrive at
∂sν 1 ρ
=− ln
∂ρν βν αν 3
and
Z ρ
1 ρ ρ ρ
sν = − dρ ln 3 = − [ ln 3 − 1 ]
0 βν αν βν αν
and
Eν Eν
sν = − [ ln( ) − 1]
βν αV ν 3 dν
Eν V
Sν (V ) − Sν (V0 ) = ln
βν V0
or
V βνkEν
∆S = k ln( )
V0
Here β = h/k and
V Eν
∆S = k ln( ) hν ,
V0
which completes this outline of this great, conceptually impressive derivation, which focussed on the
quantum nature of the radiation as opposed to Planck’s focus on the wall oscillators.

1.8.2 Einstein’s Derivation of ρ(ν, T ) - Detailed Balance


m

Einstein’s objections to Planck’s mixed derivation, which involved continuity for the fields but
discreteness for the oscillators, led him to develop a more general derivation of the black-body
radiation law based on thermodynamics and the existence of stationary states. He used the following
detailed–balance method.
Proof:
m Ref: Kittel or Bohm. & Phys. Zs. 18, 121 (1917).
CHAPTER 1. HISTORICAL AND EXPERIMENTAL ORIGINS 14

Consider two levels of the wall oscillators or “atoms,” with


N1 = # atoms in stationary state 1
N2 = # atoms in stationary state 2

—————— 2
↑ ↓
—————— 1
Transitions or energy transfers occur between these two levels and in equilibrium at a temperature
T, the ratio of these occupation numbers is

N2 e−E2 /kT
= −E /kT = e−hν/kT ≡ X.
N1 e 1
Now consider the number of excitations from 1 → 2 per sec:
dN1→2
= B1→2 N1 ρ(ν),
dt
and the number of transitions from 2 → 1 per sec :
dN2→1
= A2→1 N2 + B2→1 N2 ρν .
dt
Here A and B are Einstein’s transition probability coefficients, where the B terms: (B1→2 and
B2→1 ) refer to excitation and emission induced by the electromagnetic energy density ρν . and the A
term: (A2→1 ) is a new feature introduced by Einstein of spontaneous or natural emission, which is
independent of the radiation density ρν , but does depend on having some occupation of the upper
state. Spontaneous emission, as we shall see later, is due to fluctuations of quantized electromagnetic
fields. For both classes of transitions, the transition rate depends on the occupation number N1 or
N2 of the originating or initial state. For equilibrium, the transitions into and out of state 2 must
balance and Ṅ2→1 = Ṅ1→2 . Also, from the detailed balance theorem, which is derived later from
time reversal invariance, we have
B1→2 = B2→1 ≡ B.
Therefore, with A ≡ A2→1 ,
A
ρ= X + Xρ,
B
and
X A/B
ρ = A/B = hν/kT = αν 3 /(ehν/kT − 1),
1−X e −1
which is Planck’s form. Wien’s general thermodynamic result that ρ = αν 3 Φ(ν/kT ) is then used to
fix A/B as αν.

1.9 Specific Heat -Solids


Another example of the failure of classical theory, and the need to quantize, is the behavior of the
specific heat of solids, especially at low temperatures. Let me first present a simplified version of
the classical theory and then proceed to the Einstein and then the Debye theories.

1.9.1 Classical Theory – DuLong and Petit


The specific heat at constant volume is
d
CV = U |V ,
dT
where the internal energy of the solid is
Z ∞
U= E(ν, T ) n(ν) dν
0
CHAPTER 1. HISTORICAL AND EXPERIMENTAL ORIGINS 15

Here n(ν) dν gives the # of modes of oscillation of the solid per volume, in the interval dν, for
vibrations of frequency ν. Also E(ν, T ) is the energy
mode of the solid, which we need to determine. For
the classical case, one uses the equipartition of energy, i.e. each degree of freedom receives energy
kT /2 so that E(ν, T ) = 2 kT
2 = kT where the harmonic motion described by the Hamiltonian form
H = p2 + q 2 has 2 degrees of freedom, which explains the above factor of 2. Now
Z ∞
d
CV = kT n(ν) dν = k(3N0 )
dT 0

The integral over the number of modes n(ν) equals Avagadro’s number N0 times three, to account
for Rthe three possible directions (x,y,z) of the normal modes for each atom:

0
n(ν)dν = Total # modes
vol over all frequencies = 3N0
We find therefore that
cal
CV = 3R = 6 o ,
K − mole
Joule
where R = 8.3143 o K−mole = gas constant= kN0 . This is the classical DuLong-Petit result; note
that CV is independent of T in disagreement with low temperature experimental results, which show
that CV goes to zero at low temperatures. To resolve that discrepancy, Einstein introduced the idea
of quantizing the individual “atoms” or matter oscillators.

1.9.2 Einstein’s Theory of Specific Heat


Take N0 atoms bound as a quantized harmonic oscillators, all at the same frequency. Einstein
quantized the individual atoms using Planck’s quantization of the harmonic oscillator.
These uncoupled atoms can be pictured as:

ν• ν• ν•
ν• ν• ν•
ν• ν• ν•
Using Planck’s result for the average energy of each oscillator in equilibrium at temperature T

E(ν) = ,
ehν/kT − 1
the classical result is recovered in the limit −→ kT, or equivalently at high temperature kT >> hν.
h→0
If all atoms oscillate at a single frequency νc , then the frequency distribution of the number of
modes/volume is a sharply peaked function called the Dirac delta function (see later) and

n(ν 0 ) = 3N0 δ(ν 0 − νc ),

so that Z
n(ν)dν = 3N0 .

The integral over frequency can be done and


d hνc hνc hνc
CV = [ 3N0 hν /kT ] = 3N0 hνc 2
ehνc /kT
dT e c −1 (e /kT − 1) kT 2

Now define Θ ≡ hνc /k, which is a property of particular solid in that its compressibility determines
νc . Reexpressing the specific heat in terms of Θ, we obtain Einstein’s result:

Θ 2 eΘ/T
CV = 3R ( ) .
T (eΘ/T − 1)2
Note that for T  Θ, or h̄ → 0, the above result returns to the classical Petit–DuLong result, 3
R . However, at low temperatures T << Θ, we get
Θ 2 −Θ/T
CV = 3R ( ) e −→ 0,
T
CHAPTER 1. HISTORICAL AND EXPERIMENTAL ORIGINS 16

i.e. an exponential approach to zero at low temperatures! This result agrees with Nernst’s theorem
n
since Z T Z T Z T
dQ dQ dT C(T )
∆S = = = dT,
0 T 0 dT T 0 T
for zero temperature, we use Nernst’s theorem and take S(0) = 0 and therefore examination of the
integrand shows that C(T → 0) → 0, at least as fast or faster than T! Einstein’s result of exponential
dependence on T result satisfies this general condition, but it does not agree with experiment! The
specific heat CV is observed to approach zero as a positive power of T.

1.9.3 Debye’s Theory of Specific Heat


Debye improved Einstein’s theory by coupling the atoms and quantizing the elastic normal modes
of the whole system, instead of quantizing the motion of the individual atoms. As a result, Debye’s
theory yielded correct low temperature behavior. In Debye’s quantization of the normal modes of
the whole system, we find the early ideas of quantized field theory o .
Simplified Version of Debye’s Theory
The average energy of the quantized modes of the solid’s vibration are taken to have the average
energy
E(ν, T ) = hν/(ehν/kT − 1).
Now the number of modes is described by

n(ν)dν 1 d~k 3 1 ν 2 dν
=3 = 4πk 2 dk = 3(4π) 3
V π 8 8 π c
which gives the number of modes per volume. The factor of 3 enters here since a solid has 2 transverse
modes plus 1 longitudinal mode. Here k = ω/c = 2πν/c refers to waves in the medium [their speed
“c” should really be given as an average of the transverse and longitudinal wave speeds, using say
3/c2 = (1/c2l ) + (2/c2t ) ].
A natural cut-off νc of the integral over all frequencies is provided by the finite number of atoms
in a unit volume:
Z νc
12π νc3
dν nν = = 3 N0 ,
0 3 c3
for one mole, which roughly fixes νc . We define a Debye temperature ,Θ ≡ hνc /k, which is charac-
teristic of the particular material and therefore should be determined by correct treatment of elastic
medium properties. The internal energy of the solid is (with x ≡ hν/kT )
νc Θ/T
ν 2 dν (kT )x 12π(kT )3 x2
Z Z
U= E(ν, T )12π 3 = dx
0 c 0 ex − 1 (hc)3
Θ/T
12π(kT )4 x3 dx 12π νc 3 (kT )4
Z Z
U= = ( ) 3 ···
(hc)3 0 ex − 1 3 c (hνc )3
Z Θ/T 3
T4 T4
Z
x dx
U = (3N0 k) 3 = 3R 3 ···.
Θ3 0 ex − 1 Θ3
Therefore
1/ξ
x3 dx
Z
3/ξ
CV = 3R{12ξ 3 x
− 1/ξ } ≡ 3R D(T /Θ),
0 e −1 e −1
with ξ ≡ T /Θ. At low temperature ,

12π 4 R 3
CV (T → 0) → T ,
5 Θ3
n Nernst’s Theorem (3rd Law of Thermodynamics) states that the entropy of every system at absolute zero can

always be taken equal to zero. (E. Fermi, Thermodynamics p. 142 )


o Indeed, such quantized modes of a solid are called “phonons” in analogy to the photons of the quantized electro-

magnetic field.
CHAPTER 1. HISTORICAL AND EXPERIMENTAL ORIGINS 17

while the classical result is obtained at higher temperatures CV (T  Θ) → 3R.


Note that this prove is a simplified version. The difficulty of determining of νc or Θ is avoided
here; it is the subject of solid state physics. Also, the density of states has been simplified; we
consider no bands, no impurities, and do not treat the longitudinal and transverse motion correctly,
etc. However, the above version does not yield an exponential, but rather a power law fall off of CV
with decreasing T, as observed.

1.10 Photoelectric and Compton Effects


I will only state the general conclusions deduced by Einstein from his theory of the photoelectric
effect. On the basis of the photon (or quanta) theory of light, Einstein explained the photoelectric
effect; his theory received its generally accepted experimental confirmation by xx in 19xx. According
to the photon theory light comes in energy bundles E = hν = h̄ω (and with momentum p~ = h̄ ~k = Ec k̂
), which it could exchange with electrons in the manner of a particle.
Using conservation of energy
1 1
h̄ω = mv 2 + φ = mv 2 + h̄ωc
2 2
(conservation of momentum taken care of by metal body of large mass)
φ = h̄ωc = work function.
The following facts were immediately explained:

1. The light intensity (∝ amplitude2 or ∝ # photons/sec) effects the current (electrons/sec),


but not the kinetic energy of the electrons.
2. The kinetic energy is linearly proportional to the frequency ν, i.e. the kinetic energy depends
on the color, with the slope yielding h.
3. Each material has a characteristic work function, φ, below which a given color produces no
electrons.
4. A very short time delay occurs between the impact on the surface and emission of the electron
(classical theory would require a very long time for energy to accumulate of the order of
seconds). (The classical picture also would have the kinetic energy depend on the light intensity
which is proportional to E 2 + H 2 ).

1.11 Compton Effect


The Compton effect [∼ 1924, PR 25 (1925) 306] is closely related to the photoelectric effect, except
for dealing with X-rays (whose wave length is of atomic size ∼ 10−8 cm) and with free or weakly
bound electrons. It demonstrated that the discrete packets of light energy (photons) not only carried
energy Eγ = h̄ω, but also momentum p~γ = h̄~k = Ec k̂. The result (see appendix) is that, independent
of scattering material and of incident λ, the change ∆λ is

∆λ = 4π (sin2 θ/2) (max at θ = π)
mc
The Compton wave length of the electron is λ –e = h̄/mc = 3.86 × 10−11 cm = 386 F.
On the other hand, classical theory predicted that, as a consequence of the electron being accel-
erated by the incident wave, it would radiate at a shifted wave length ∆λ ∝ λ. A full classical result
(see Messiah p. 15) is simply a Doppler shift:
pc
∆λ = 2λ sin2 θ/2.
E − pc
The Compton experiment involved passing X-rays through supersaturated water vapor. Primary
electrons (the recoil ones) and secondary electrons (originating from second acatterings) were seen.
As a result of these experiments, it was concluded that light occurred as photons with

Eγ = h̄ω;
CHAPTER 1. HISTORICAL AND EXPERIMENTAL ORIGINS 18

p~γ = h̄~k,
and ω = kc. The left hand side of these relations has particle, the right side wave–like attributes.
The great problem was then to resolve this wave-particle paradox for photons. However, the paradox
was first expanded to the case of matter, by Prince Louis DeBroglie.
Appendix
Compton Effect–Derivation

mc2 + h̄ω = h̄ω 0 + E 0


h̄k 0 cos θ + p0 cos φ = h̄k
h̄k 0 sin θ = p0 sin φ
p0 2 = h̄2 k 02 sin2 θ + h̄2 (k 0 cos θ − k)2
p02 = h̄2 (k 02 + k 2 − 2kk 0 cos θ)
= (mc2 + h̄ω − h̄ω 0 )2 − m2 c4
= h̄2 ω 2 + h̄2 ω 02 − 2h̄2 ωω 0 + m2 c4 + 2mc2 h̄(ω − ω 0 ) − m2 c4
2ωω 0 h̄2 (cos θ − 1) = 2mc2 h̄(ω 0 − ω)
2π 2π ω − ω0 2h̄
∆λ = λ0 − λ = c( − ) = 2πc = 2πc 2 sin2 θ/2
ω0 ω ωω 0 mc

∆λ = 4π( )(sin θ/2)2
mc
CHAPTER 1. HISTORICAL AND EXPERIMENTAL ORIGINS 19

1.12 The Atomic System


Let me continue our survey of the empirical basis for the failure of classical theory by discussing the
development of the atomic or nuclear model. As determined by J. J. Thomson the electron has a
well defined charge and mass, it moves in a “ray” or beam - originally call β-rays. The existence of
these trajectories and furthermore the direct observations of electron paths by Wilson in 1911 were
the main reasons for viewing these β-rays as “particles.” (If the diffraction properties of electrons
had been observed before these particle aspects, we probably would tend to view the electron as a
wave and have a psychological difficulty picturing them as particles. I mention this as a step towards
the therapy that will be required later - i.e., of freeing ourselves from the particle–wave syndrome.)
Based on the view of the electron as a particle, the Rutherford nuclear model (based on the
scattering of alpha particles from atoms) was used by Bohr to explain atomic spectra, mostly of
simple atoms such as H, He+ , Li++ , etc. That is a familiar story to you; nevertheless, for later use
I would like to highlight a few aspects of the Rutherford–Bohr nuclear model.
The first point is that the Rutherford model led to a striking failure of classical theory, which
would predict the death of a Rutherford atom in about ∼ 10−10 sec. (See attached estimate).
Clearly, the obvious stability of our universe (on that time scale) shows that direct application of
classical radiation theory is wrong, if the Rutherford model is correct.
Classical theory also would yield a continuous spectrum.

Stability of Matter
The hydrogen atom would collapse in 10−9 to 10−10 sec.

2 e2 2
Ė = − a
3 c3
(time averaged) energy loss.

2 e2 v 2 2 4 e2 1 2 1
=− ( ) = − ω ( mv 2 )
3 c3 r 3 mc2 c 2
T = −E
E = T + V = T − 2T = −T Virial Theorem :
2
4 ω 4 ω2
Ė = + r0 E = − r0 |E|
3 c 3 c
E = E0 e−λ0 t

4 ω2 4 4π 2 4.5 × 10−1
λ0 = r0 = (2.8 × 10−13 ) 2 (3 × 1010 )2 =
3 c 3 λ λ2
e2 h̄ 1 –
r0 = = 2.8 × 10−13 cm = α = λelectron
mc2 mc 137

ω = 2πf = c
λ
c = fλ
for visible light
λ = 1500 Å = 1.5 × 10−5 cm

.45 4.5 1
λ0 = −5 2
= 2
× 109 '
(1.5 × 10 ) (1.5) τ1/2

τ1/2 = .5 × 10−9 = 5 × 10−10 sec !


If you try to balance with absorption of external radiation then the spectral description of external
radiation (black–body radiation) is wrong. We need a basic theory to explain the stability of matter!
CHAPTER 1. HISTORICAL AND EXPERIMENTAL ORIGINS 20

1.13 Bohr’s Atom - Postulates (1913) – Direct quotes


I “An atomic system possesses a number of states in which no emission of energy radiation takes
place, even if the particles are in motion relative to each other, and such an emission is to be
expected in ordinary electrodynamics. The states are denoted as the “stationary” states of
the system under consideration.”
II “The dynamical equilibrium of the systems in the stationary states is governed by the ordinary
laws of mechanics, while these laws do not hold for the transition from one stationary state to
another.”
III “Any emission or absorption of energy radiation will correspond to the transition between two
stationary states. The radiation emitted during such a transition is homogeneous and the
frequency ν is determined by the relation

hν = E1 − E2 ,

where h is Planck’s constant and E1 and E2 are the energies of the system in the two stationary
states”.
IV “The various possible stationary states of a system consisting of an electron rotating around a
positive nucleus are determined by the relation T = nh̄ω/2, where T is the mean value of the
kinetic energy of the system, ω the (angular) frequency of rotation, and n a whole number”.

Note:
mv 2
= nh̄ω/2; mv 2 = nh̄ω; mvr = nh̄ = `
2
Bohr’s Atom

Ze2 mv 2
ma = 2
= = mrω 2 ,
r r
Here ω = frequency of rotation (classical mechanics; use postulate II.
Quantization; use postulate IV.

` = mvr = mr2 ω = nh̄


or
I I
p dq = ` dθ = 2π` = nh

(Bohr-Sommerfeld quantization – for circular orbits)

mr3 ω 2 = Ze2
m2 r4 ω 2 = n2 h̄2
n2
rn = a0
Z
some useful relations
h̄2
a0 = = Bohr radius = .53 Å
me2
h̄ h̄c 1 –
a0 = 2
= λ e = .53 Å
mc e α

–e = h̄ = h̄c = 3.86 × 10−11 cm


λ
mc mc2

e2 1
α= =
h̄c 137
CHAPTER 1. HISTORICAL AND EXPERIMENTAL ORIGINS 21

e2 –e = 2.8 fm
r0 = classical electron radius = =αλ
mc2

–e mc2 = 197 MeV − F


h̄c = λ

mc2 = .51 MeV

1 Ze2 1 Ze2
En = mv 2 − =−
2 rn 2 rn

En 1 Z 2 e2 1 Z 2α 1 2πRc
= ωn = − = − c = − 2 Z2
h̄ 2 h̄a0 n2 2a0 n2 n
Rc
(νn = − )
n2
also, for Z = 1 :
α α2
E1 = − h̄c = − mc2 = −13.6 eV
2a0 2
1
1 Z 2α 1
Rc = c= 137
3 × 1010 sec−1 = 3.29 × 1015 sec−1
4π a0 4π .53 × 10−8
1 1
ωnm = −2πRc ( − 2) > 0 m < n
n2 m
Note ω is angular frequency of rotation in orbit – for a classical system this would also be the
frequency of the emitted radiation. But in Bohr’s model the radiation frequency is given by p

ωnm = ωn − ωm m < n.

2πRc
ωn = −
n2
The classical result should be obtained in the “correspondence” region; i.e. for high n’s experi-
ment showed radiation of “classical” frequency.
1 1
ωn+1,n = −2πRc ( − 2)
(n + 1)2 n
2n + 1
= 2πRc
(n + 1)2 n2
for n → very large n >> 1
2πRc“2”
ωn+1,n =
n3
extra factor of “2” important! Basic reason for Bohr’s use of T = nh̄ω/“2”.
Classical frequency in the nth orbit is:

nh̄ h̄Z 2 1 Z 2 αc 1
ωc` = 2
= 3
=
mr ma0 a0 n a0 n3
4πRc 1 dE
ωc` = = lim ωn+1,n = |n→∞ ,
n3 n→∞ h̄ dn
which is an example of the correspondence principle: for high quantum numbers get classical results!
Served historically to deduce Bohr–Sommerfeld quantization and as guide to the “new” quantum
mechanics!
p Lyman ( m=1), Balmer (m=2), Pashen (m=3), Brackett (m=4), and Pfund (m=5) are the names of the spectral

series.
CHAPTER 1. HISTORICAL AND EXPERIMENTAL ORIGINS 22

1.13.1 Origins of Matrix Mechanics


Another useful guide to the formulation of the “new quantum mechanics” was provided by the
frequency addition rules for observed spectral lines. The experimental results showed that the
Rydberg–Ritz combination principle held for atomic systems, i.e.,

ωnm = En − Em

ωnk + ωkm = ωnm


According to classical physics the frequencies of radiation are determined by the frequencies
involved in the periodic motion of the electron’s trajectory. For multiply periodic systems ( with
f ≤ 3N degrees of freedom)

X
~q = ~q (ni · · · nf ) ei(n1 ω1 +n2 ω2 +···nf ωf )t
0
amplitudes give intensity of radiation; here, ωi · · · ωf are the classical fundamental frequencies. We
observe the frequencies n1 ω1 + n2 ω2 + ·nf ωf = ω , which is the classical rule for the observed fre-
quencies. It is different from the Rydberg– Ritz combination principle and the relationship between
these rules based on the correspondence principle (i.e. that they should agree for large quantum
numbers) is the clue that W. Heisenberg used to deduce a new dynamics. After deducing the rules
for these coefficients, he and Max Born found out the rules were identical to those for non-commuting
matrices, at which point Matrix mechanics was born.

1.13.2 Additional Support for Bohr’s Stationary States


Let me stop this train of thought and return to itemizing the basis for the breakdown of the classical
and rise of the quantum theory.
At this point, I just mention two experimental results, independent of the spectral observations,
which give support to Bohr’s postulate of the existence of stationary states En . These are
1. Moseley’s Law (1913) for X-rays, showed that atomic number Z can be associated with charge
on nucleus – and was consistent with Bohr’s theory.
2. Franck–Hertz (1914)
Without knowing of Bohr’s work, F–H observed the discrete excitation of atoms, undergoing
collisions with electrons – they established the existence of the stationary states En - directly
– (not just En − Em ).
3. Stern–Gerlach (1922) - support for idea of stationary states.

1.14 Failures of Bohr’s Theory


Despite the great success of the Bohr postulates, they formed a mixed bag of classical plus quantiza-
tion rules, which was conceptually unsatisfying, especially to Bohr himself. In addition, the theory
failed in several important regards; namely;
1. The intensity (and forbiddance) of spectral lines could not be predicted (a clever correspon-
dence principle argument gave some guidance) – Transitions.
2. The extension to the He atom (and heavier) went beyond the multiply-periodic restrictions of
the theory.
1
3. Diatomic molecule spectra had unusual 2 integer quantum numbers of unexplained origin.
4. Couldn’t be used for nonperiodic continuum, scattering problems.
5. anomalous Zeeman and 2nd order Stark effects could not be handled properly by old QM A
better theory was clearly needed and the “new” q.m. succeeded in all of the above cases.
q

q Rotator – Molecular Spectra


CHAPTER 1. HISTORICAL AND EXPERIMENTAL ORIGINS 23

Physical Constants
c = 3 × 1010 cm/sec

e = 4.8 × 10−10 esu = 1.6 × 10−19 coul

1 MeV = 1.6 × 10−13 joules

h̄ = 6.58 × 10−22 MeV − sec = 1.0545 × 10−27 erg − sec

k = 8.6 × 10−11 MeV/deg

me c2 = .51 MeV

mπ c2 ∼
= 140 MeV

Mp c2 = 938.3 MeV

α = e2 /h̄c = 1/137.04

–e = h̄ = 3.86 × 10−13 m = 3.86 × 10−11 cm = 3.86 × 10−3 Å = 386F.


λ
me c

e2 –e = 2.8F = α2 a0
re = =αλ
mc2

h̄c = .51(3.86 × 10−11 )MeV − cm = 197 MeV − F = λ


–e mc2

h̄2 1 – –
a0 = 2
= λ e = 137 λe = .53 Å
me α

α –e
α2 λ α2
E1 = − h̄c = − – mc2 = − mc2 = 13.6 eV
2a0 2 λe 2

e2 h̄2 1 2
=− =− ( )
2a0 2m a0
Applying Bohr–Sommerfeld quantization to a rigid rotor model for molecular rotations we have
Z
` dθ = nh = 2πIω

or I
I = nh̄.

Therefore, the energy of rotation is given in the old quantum theory by


1 2 1 2 2
E= Iw = n h̄ .
2 2I
The new quantum theory yields energies E ∝ (n + 21 )2 − 1
4
no 1/2 quantum numbers needed to fit data.
CHAPTER 1. HISTORICAL AND EXPERIMENTAL ORIGINS 24

1.14.1 Electron Waves


Light exhibits wave properties in interference and diffraction experiments, but it reveals particle
aspects in the photoelectric, Compton, and (according to Einstein) in black-body radiation effects.
This paradoxical situation was further complicated by the suggestion of Prince Louis de Broglie’s
thesis (which was examined by Einstein) that electrons could also display wave-like properties.
Earlier evidence had indicated the electron was a “particle.” It had well defined values of e/m
and moved in well defined paths as beams or rays. This was determined by J. J Thomson, who
is credited with discovering the electron. Furthermore, Wilson’s cloud chamber (1911) provided
an almost direct view of the electron‘s trajectory. That fact, along with the success of the Bohr-
Rutherford nuclear model, seemed to give firm support to the particle aspect of the electron.
Nevertheless, based on an analogy with the photon and using the notion of a group velocity as
the “particle’s” velocity, de Broglie deduced that a wave length of
h h
λ= =√ ,
p 2mE
or
h
p=
= h̄k,
λ
could be associated with an electron (or any other mass) with momentum p. That result is analogous
to the photon case for which E/c = p = h̄ω/c = h̄k = hλ, but differs significantly in attributing
this wave length to matter waves. Exactly, what these waves signified was a very mystical part of
de Broglie’s work.
Motivated by an unrelated patent fight in the U.S., Davisson and Germer (1927), and indepen-
dently G.P. Thomson(1927)r , discovered that electrons of 1-100 keV displayed diffraction patterns
when incident upon crystals of various kinds. (Later verified for atomic and nuclear beams).
Thus we arrive at the full wave-particle duality or wave-particle paradox. Both light and matter
display wave (field) and particle aspects depending on the experiment involved. This represent a
full breakdown of our classical assumption concerning fields (spread-out) and particles (localized,
carriers of energy E and momentum p~). The fundamental assumptions of classical theory, such as
the separation of object and observer, come next into question.

1.14.2 Resolution of the Wave-Particle Paradox


This paradox is resolved by simply accepting it – that has caused some problem among logicians,
who worry about the excluded middle, etc. Nature is telling us something, and we are obliged to
accept it. What we are told is called by Bohr - duality or complementarity
Sometimes an experiment measures the wave-like aspects, other experiments the particle - like
aspects, of a system. Depending on the experiment, one sees one aspect or the other.
To understand how such an unusual statement can make sense, we have to address the question
that is most basic to science – How do we measure?

1.15 The Measurement Process - Uncertainty Principle


R
For small systems, i.e., systems of small action, L dt, the classical concept of measurement becomes
invalid. That fact is the most basic idea of quantum theory. Earlier, we discussed the basic classical
assumption that the effect of the measuring device could be made not to appreciably alter the object
measured, or at least we could precisely correct for the act of measurement using known laws of
dynamics. Now we must reconsider that assumption in view of the small action (as scaled by h̄) and
the limitations imposed by the discrete nature of photons and of atomic systems and of the wave
nature of electrons.
We shall study the basic process of measurement for “small” systems by using gedanken ex-
periments, (as adopted by Heisenberg from Einstein) to gain physical insight into the meaning of
r The son of the man who established the electron as a “particle.”
CHAPTER 1. HISTORICAL AND EXPERIMENTAL ORIGINS 25

the uncertainty principle. Later we shall develop a formalism for expressing these basic consider-
ations (we are reversing the historical order at this stage) s . We shall learn from these gedanken
experiments that:

1. It is impossible to measure both the position and momentum of a “particle” simultaneously


with infinite precision. The limit placed on our ability to measure is expressed as the uncer-
tainty principle ∆px ∆x ≥ h̄.
2. Any any attempt to determine the wave-like aspects (distributed in space - i.e., ∆px small,
∆x large) of a small system destroys the possibility of determining the particle-like aspects
(localized so that ∆x is small and ∆px is large) of a small system. This expresses the comple-
mentary nature of the wave and particle pictures. They serve only as useful guides to thinking
which are not mutually exclusive, but apply to experiments in which one asks complementary
questions.

3. A probabilistic approach is required due to our ability to measure being limited. The probability
is attributed not just to an ensemble of electrons, but to each individual electron (or photon).
(We shall describe this probability in terms of a probability amplitude or wave function Ψ,
which is to be defined not as a wave in 3-D in ordinary space, but in the 3N dimensional
abstract space - configuration space of the N particles) t .

Let us consider some of the gedanken experiments.

s As described in Heisenberg’s book and in his lectures, the formalism was developed first and then the interpretation

was developed. His original approach was to deal directly and only with observable quantities, the transitions between
levels and the (dipole) radiation. He subjected the dipole moment eq to the conditions that they yield the Rydberg-
Ritz rules and satisfy the correspondence principle by yielding classical radiation rules in the classical limit. The
dipole quantities (now called dipole matrix elements) were labelled by the quantum numbers of the 2 associated levels
using an extension of the Fourier decomposition of a multi-periodic motion.
t Phase space refers to the 2f space of both momentum p and location q, each of which has f degrees of freedom
CHAPTER 1. HISTORICAL AND EXPERIMENTAL ORIGINS 26

1.15.1 The Uncertainty Principle - Heisenberg’s Microscope


Heisenberg used the following hypothetical microscope as a means of illustrating the limitation
nature imposes on our ability to simultaneously measure an electron’s position and momentum,
i.e., ∆px ∆x ≥ h̄. Here ∆px and ∆x are the uncertainties in measuring a system’s momentum
and location simultaneously. Historically, the above relation was obtained mathematically from
the operators in [p, q] = h̄/i; it was later that the full meaning of that relationship was revealed
by gedanken experiments such as the following microscope case. To measure the position of an

AKA Plate
 A
 A
 A
 A
 ................................................ A
....... . .............
..... ............... ............
A....
AKA ...............................................  Lens
A 
A 
A 
A 2θ 
-
-
- A 
-
- Aq
-
- Electron px0
h
Light pγ = λ0

Figure 1.2: Heisenberg’s Microscope

electron one needs to define the procedures. Heisenberg emphasizes that all physical arguments deal
with observables, which are made meaningful by stipulating the detection procedures. Hence, the
microscope is to have a plate, a lens, an electron in the field of view with initial momentum px0
(known precisely). Light in the form of X-rays or γ-rays is incident from the left with known wave
length λ0 u . We know v from optics that the resolving power of a lens is given by
λ0 1 λ0 1
∆x = ' .
2 sin θ 2 sin θ
Hence, we should use a low λ0 (say λ0 ∼ 10−9 corresponding to atomic dimensions and Eγ '
.12 MeV) and a large (θ large) lens for an optimum small value of ∆x. In fact, for small λ and large
θ, we could get ∆x down to any required accuracy and hence, use the microscope to measure the
electron’s position very accurately ( see later for a lower limit to ∆x). However, as we do this the
energy of the incident photon increases, which can deliver an uncontrollable and increasing amount
of momentum to the electron.
Of course, to minimize the effect of the light on the electron, one photon should be considered.
Thus, the γ + e → γ 0 + e0 collision should be described by the Compton effect equations. Instead,
let us consider the two cases shown in the next figure, corresponding to limiting recoil photons that
are detected.
Using conservation of momentum for the two angles ±θ, we have
h h
+ px0 = p0x + sin θ
λ0 λ1
h h
+ px0 = px − sin θ.
λ0 λ2
We could now use conservation of energy (and the Compton effect algebra), instead we note that
for high energy γ’s that λ1 ' λ2 ' λ0 and hence the momentum spread is
u Note this example uses corpuscular and wave theories, i.e., photon and lens, which is typical since measurements

are always macroscopic and hence involve classical concepts in part.


v See Peierl’s book for an amusing story about Heisenberg, who had trouble on his Ph. D. oral exam with this

formula.
CHAPTER 1. HISTORICAL AND EXPERIMENTAL ORIGINS 27

...................................... ......................................
.................. ........... .................. ...........
.................... .. .................... ..
.... ............ Lens .... ............ Lens
K ................................................ 
A
A   AK ................................................ 
A 
A  A 
A  A 
A  A 
h  Photon h
- Photon
A
λ2 - A
λ1
- AA -
- KA  - - A 
-
- Aq -
- q

AH -
Electron px0 Electron px0
-
- -
-
AA
U HH
Light pγ = λh Electron p0 Light pγ = λh0
j p0
Electron
H
0

Figure 1.3: Heisenberg’s Lens–Two limiting Angles

h
p0x − px = ∆px = 2 sin θ
λ0
We could minimize ∆px by using larger λ (i.e. light) and a smaller lens ( small θ ) , but only at the
price of increasing ∆x. Put together, we have ∆px ∆x ' h.
Some additional remarks:
• Note: (see Heisenberg’s book for the following example) One idea is to make the lens moveable
and measure its recoil from the light quanta to make more certain the path of the photon
within the bundle of rays in θ! However, then one must consider how to measure the x and p
of the lens.
• Also, one could detect both the electron and read a fixed location scale attached to the recoiling
lens simultaneously. Then, we need to detect two photons, (1 from electron, 1 from scale),
then recoil of lens does not give us direction of light from the election - ad infinitum. See W.
Heisenberg. p. 22.
• If use N photons, then still no gain since ∆x = √ λ0 , but
N sin θ

√ sin θ
∆p ∝ Nh
λ0
and we get
∆x ∆px = h,
as before.
• Limitation on ∆x
At this point,it is worthwhile to realize that the measurement of ∆x is limited in itself. This
point is discussed by in Gottfried’s book and I wish to call attention to it. Namely, one can
not always measure ∆x with infinite precision, irregardless of the effect on p~.
To determine the limitation imposed on our ability to measure ∆x, recall that light of wave-
length λ0 goes into the lens and
λ0
∆x ∼
sin θ
(earlier we had set λ0 ∼ λ0 (Also ∆px ∼ h̄
λ0 sin θ so, that ∆px ∆x ' h̄.)
To measure location with zero uncertainty ∆x, we must use a very short wavelength λ0 →
0, since sin θ can be made only so large ∼ 1. As λ0 → 0, ω 0 → ∞ and Eγ → ∞. With so
much energy available, electron–positron pairs can be produced! Which electron is the original
CHAPTER 1. HISTORICAL AND EXPERIMENTAL ORIGINS 28

is therefore unknown to the order of the so called “zitterbewegung” or trembling motion.


Gottfried determines this lower limit to ∆x by noting that
2h̄ω θ
ω 0 = ω[1 + sin2 ]−1 .
mc2 2
For ω → ∞, ω 0 → 12 mc2 sin−2 θ2 . Therefore, ω 0 is bounded (except for θ = 0), and thus λ0 is
also bounded, as is ∆x. We have
c 2π 2πch̄ h̄ –e .
∆x ∼ 0
∝ 2
2 sin2 θ/2 ∝ ∝ λ
sin θ ω sin θ mc mc2
Thus, ∆x < λ –e is a limit on our ability to measure location of the electron. Also note that the
ratio of the Compton to the deBroglie wavelengths
–e
λ p v
–B = mc = c << 1
λ
expresses our limit to nonrelativistic quantum mechanics. In that limit , we require
–e << λ
λ –B

so we can use nonrelativistic QM only


until pairs can be produced. Pair production was not known until invented by Dirac and
subsequently verified by experiments.
• Limit on Times
there is also an associated limit on our ability to measure time.

∆t ≥ T 0 ,
using Compton equations this reduces to

1 h̄ 1 –
∆t ≥ >> = λe /c
ω0 mc c
–e /c limits our time measurements to the time it takes light to travel over λ
Thus ∆t ≥ λ –e !

Later we shall worry about preparation of states with precise p~ and ~x values!
CHAPTER 1. HISTORICAL AND EXPERIMENTAL ORIGINS 29

1.15.2 The Uncertainty Principle - Single Slit


We consider now another example of how to measure position, with less dependence on invoking
both the wave and corpuscular theories of light than was needed in the lens + photon example. A
single slit fixes the location of an electron which passes through it normally, as illustrated below.

..................
.................................
.. .
................ .......... .......
...... ... ...
.... .... ..
... .. ...
... ... ...
p~ . . .
∆x ..... ..... .....
-
.. .. ..
.. .. ..
.
. . ..!
.... ..... ....
.......... ........ ......
....
.. .
..............................
....................

If motion in the x–direction is precisely known we have a beam, ray, or trajectory and ∆x is
determined. Thus, we expect ∆x = d is the width of beam, but we have diffraction and as d is
decreased to order of the de Broglie wavelength, the electron displays its wave nature and diffracts
as shown above. We have
λ
sin α ∼ ;
d
and with λ = hp ,
h h λ h h
∆px = sin α = = =
λ λ d d ∆x
Thus ∆px ∆x ∼ h.
We could let the screen move – then questions of its momentum and position arise. If we let
mass M of the screen → ∞, then δxscreen → 0 and δpscreen ≥ ∆px .

1.15.3 The Double Slit-Complementarity & Probability


The double slit is a very useful way to show how the uncertainty principle leads us to the idea of
complementarity and a resolution of the wave–particle paradox. The resolution consists of giving up
microscopic causality and instead formulating
......................
..
a probabilistic approach.
.........
......
.................. .
....... .......
..... .
.............. .... .....
........................... ..... ....
........ ................. .... ....
A .................. ... ......... ....
......... ............... .... ...... ....
p~ ............ ..... ........ .......
..................... .......... ........
-
....... ........... ... .... ...
B ..................... ..... .......... ....
......... ............... ... ....
............................ .... ...
............ . .
.... ..
. . ...
. .... ......
................. ....
...
.. . ....
.... ........
...............

Figure 1.4: Double Slit with Screen

The double slit or Young experiment is, of course, the traditional means of demonstrating inter-
ference of waves - the waves pass through both slits maintaining their phase (coherence) and maxima
and minima are determined by their phases at points on a screen. We now consider the case of a
single electron (or photon) incident on such a set-up. (In reality one deals with a beam of electrons
on a crystal latice, or a diffraction grating).
We shall see that we can use this experiment to ask:
CHAPTER 1. HISTORICAL AND EXPERIMENTAL ORIGINS 30

1. is an electron is a wave?
and get the answer: yes - when both slits are open and no precise determination of trajectory
is made.
2. We can also ask is the electron a particle?; i.e. determine its trajectory
and get an answer: yes - by determining the hole it passed through, but in the process
destroying the interference pattern.
The two questions and the corresponding experiments are therefore complementary, but mutually
exclusive. Asking one of these questions involves steps that destroy the possibility of simultaneously
getting an answer to the other question. Let us see how this argument is made
First, we must understand the interference based on a single γ or electron. It is possible to impinge
just one γ (by lowering the intensity at a fixed frequency) or just one electron at a time – say each
second. Nevertheless, after a wait the interference pattern builds up. Therefore, the situation is not
that two γ’s (or electrons) are involved - one passing through each slit and annihilating at the dark
spots - this would be objectionable anyway because of conservation of energy. Hence, we consider a
single projectile.
To get interference, clearly both slits are involved; if one slit is covered the interference pattern
is destroyed. To answer the question: is the electron a wave? we leave the slits open - don’t interfere
and see interference - answer is yes! Question: which slit did it pass through? Answer: we didn’t
make a measurement to determine the trajectory ... we don’t know in this case!
Next question: Is the electron a particle? Let’s determine the trajectory! To determine the
trajectory, we could now consider one of Einstein’s objections (that Bohr used to clarify his viewpoint
of complementarity) .
CHAPTER 1. HISTORICAL AND EXPERIMENTAL ORIGINS 31

Alternate Discussion
One could introduce a device to measure presence of electron - say a shutter or scintillator at slit A.
If we get response in that detector, then an exchange of momentum takes place at A

∆p = 2p sin θ/2 = px θ

A
p~ -
∆s
B

Figure 1.5: Double Slit with Detector at Slit A

The uncertainty in location of the trajectory for that part passing through A is

∆x ≥
px θ

Since θ ∼ a/D,

∆x ≥ D
px a
or
–D
∆x ≥ λ
a
∆x ≥ ∆s,
where ∆s is the separation between the interference bands. Thus, if trajectory is well known the
interference pattern gets washed out. Our determination of the trajectory, by ascertaining that slit
A was involved has led to wiping out the interference effect!

Einstein’s Objection to ∆x ∆px ≥ h̄


Ref: Cropper and Bohr’s article in Einstein vol.
Einstein’s suggestion for simultaneous measuring ∆x with sufficient precision to know which slit
the electron goes through while still maintaining interference is as follows:
One measures ∆px from the recoil of a hanging double slit setup. The recoil direction tells us
which slit the electron goes through, ... ∆x is small i.e. << a and still can detect possible interference
pattern on the screen. Bohr’s reply was as in previous section, that the measurement of
h
∆px = 2p sin θ/2 ∼ p θ = θ,
λ
which implies that
h hλ
∆x ∼ = .
∆px hθ
Since
a
θ∼ ,
D
we have
λ
∆x ∼ D
a
CHAPTER 1. HISTORICAL AND EXPERIMENTAL ORIGINS 32

which means ∆x exceeds the distance between bands in the associated interference pattern. If we
determine ∆px precisely, in that act we loose knowledge of ∆x enough to destroy the interference
pattern. If the trajectory is known (particle aspect), we destroy the wave (interference) property -
complementarity again!
Any attempt that pins down trajectory enough to decide which hole was passed through alters
the system enough to destroy the interference pattern. Only get interference, when trajectory is
indeterminant.

1.15.4 Complementarity and Probability


The conclusion is that our ability to measure and hence to know is limited and we must accept
that fact in building our formalism. We have seen that even for a single electron or photon, we can
not determine a trajectory without destroying the interference pattern. The interference pattern
can be produced only if both slits are involved. Also an electron (or a γ) can not be split “a
fraction of a photon is never observed” - Dirac page 2. Without measuring the hole used and hence
determining the trajectory, we really do not know what happened between the source and the screen.
Hence, we do not make claims other than involving a probability. We give up a purely deterministic
causal viewpoint and invoke a statistical or probability amplitude Ψ(x). The probability amplitude is
introduced so it can incorporate interference in the usual way obtained for waves. This probability
amplitude concerns a single γ or electron. More about this later.
We also are led to accept the Complementarity Principle as stated by Bohr as:
” Consequently, evidence obtained under different experimental conditions cannot be compre-
hended within a single picture, but must be regarded as complementary in the sense that only the
totality of the phenomena exhausts the possible information about the objects.” ( Albert Einstein:
Philosopher-Scientist, Tudor publishing Company page 210)
or by Gottfried as:
“An experimental arrangement designed to manifest one of the classical attributes (e.g., wave
or particle-like aspects) precludes the possibility of observing at least some of the other classical
attributes.”

1.15.5 Counter Examples and Other Interpretations


Ref: 1. Bohr in Einstein vol 2. Heisenberg in Bohr vol - a partisan survey 3. Bohm
To be fair the various viewpoints concerning the interpretation of quantum mechanics should
be discussed. We will not be able to do that, it takes a lifetime. Let me just point out that these
objections are often in the interpretation or philosophy but have no effect on the actual numerical
results or formalism. They largely belong to the philosophers. Other objections are attempts to
disprove the validity of the uncertainty principle. In this regard, Einstein has provided the most
clever and challenging situations. The “Copenhagen” answer; w i.e., Bohr’s reply is given in the
above references 1 and 2. Let me present: (a) a sketch of Einstein’s objection to ∆E ∆t ≥ h̄, and
then turn to another case which Einstein also presented, which was resolved by Heisenberg; i.e., (b)
an apparent violation of the uncertainty principle.
Later in the course when we get to spin, I will discuss the Einstein, Podolsky, Rosen paradox
(EPR)(Physical Review, 47 ,777-80 (1935) and Bell’s inequalities.

Einstein’s Clock in a Box


a) First consider Einstein’s Counter example to ∆E ∆t ≥ h̄
Ref: Cropper or Bohm in Einstein vol.,Solvay Conference (1930)
The device is shown in the figure.
At a predetermined precise time, the clock moves the shutter and a γ comes out with ∆m =
E/c2 = h̄ω/c2 , causing a change ∆m in mass of the box. The ∆mg is measured by the ∆x of the
ruler attached to the box - i.e. using the balance. Thus we can get ∆t → 0 by presetting the clock.
Also
∆x → ∆m → ∆E → small
by precise measure of ∆x.
w The Copenhagen school ≡ statistical version.
CHAPTER 1. HISTORICAL AND EXPERIMENTAL ORIGINS 33

t
@ Spring
@
Radiation γ


Pointer Clock
# - Escaped Photon h̄ω
P PP
 .
......
...............
"! ............Shutter

Figure 1.6: Einstein’s Clock in a Box

... ∆E ∆t → 0 not h̄.


Bohr’s reply: After release of the γ, the box has a change in weight of g ∆m which for equilibrium
equals the force exerted on the box by the spring = g ∆m. That force acts on the box for a time t,
which is the time needed for the pointer and box to get to its new equilibrium point. An impulse
F t = g∆mt = acts on box during this time. It is this impulse we wish to determine (so we can
deduce ∆m). To measure the impulse g ∆m t, we must be sure that the act of measuring ∆x does
not impose a momentum on to the box that is big enough to make g ∆ m t unobservable. That is,
we want the ∆p imposed on the box by our observation of ∆x to be

∆p ≤ g ∆ m t
or

h̄ h̄ h̄c2 1 1 h̄∆x
∆x ≥ ≥ = = h̄ gt =
∆p gt∆m g t ∆E c2
∆E ∆E ∆t
In the last step Bohr used general relativity, wherein a change of ∆x of a clock’s position changes
its rate by ∆t = cgt2 ∆x
... ∆E ∆t ≥ h̄. QED
postscript: Later Einstein suggested subsequently opening the box - which involve two experi-
ments not one!

Two Experiments and ∆x ∆p ≥ h̄!


Suppose we know (have measured) the px of a free electron at t1 , its location is unknown. At t2 we
p
~
measure its location x2 . Using x1 = x2 + m (t1 − t2 ), we can ascertain where the electron was at the
past time t1 . Therefore at time t1 , we know
∆px → small
∆x → small.
and thus
∆px ∆x < h̄!
The above seems to violate the uncertainty principle.
However, that determination is made after the second experiment is performed at t2 . The
uncertainty principle involves only one experiment at one time and not a projection back in time!
Of course, the reverse case - measure x then p is the same situation - as is Einstein’s later opening
of the box in his clock example.
The best way for me to convince you is to use the authority of Heisenberg who stated in his
book:
CHAPTER 1. HISTORICAL AND EXPERIMENTAL ORIGINS 34

“The uncertainty principle refers to the degree of indeterminateness in the possible


present knowledge of the simultaneous values of various quantities with which the quan-
tum theory deals; it does not restrict, for example, the exactness of a position measure-
ment alone or a velocity measurement alone. Thus suppose that the velocity of a free
electron is precisely known, while the position is completely unknown. Then the prin-
ciple states that every subsequent observation of the position will alter the momentum
by an unknown and undeterminable amount such that after carrying out the experiment
our knowledge of the electronic motion is restricted by the uncertainty relation. This
may be expressed in concise and general terms by saying that every experiment destroys
some of the knowledge of the system which was obtained by previous experiments. This
formulation makes it clear that the uncertainty relation does not refer to the past; if the
velocity of the electron is at first known and the position then exactly measured, the
position for times previous to the measurement may be calculated. Then for these past
times ∆p ∆q is smaller than the usual limiting value, but this knowledge of the past
is of a purely speculative character, since it can never (because of the unknown change
in momentum caused by the position measurement) be used as an initial condition in
any calculation of the future progress of the electron and thus cannot be subjected to
experimental verification. It is a matter of personal belief whether such a calculation
concerning the past history of the electron can be ascribed any physical reality or not.”

1.15.6 Illustrations of the Uncertainty Relations


Note: Other tough cases are:
See electron here at point A on screen - know no electron at B - violates relatively - transmittal
at above the speed c! No, since Ψ is not a real material wave, not a signal it’s a probability!
Wilson Cloud Chamber shows localized trajectories See Heisenberg’s struggle on this very sig-
nificant issue, which is ultimately resolved by quantum theory of scattering!
A consequence of the prior case of measuring at two different times is that we need initial
conditions - prepare an initial state - see consequences to do physics; we have an initial value
problem, with the state Ψ stipulated at an initial time.
Above results are not reproducible but stochastic. ... To measure full aspects (i.e. Ψ) of system
many cases are needed - repeating experiment needed to define ∆px - Although Ψ applies to each
electron - we need numerous measurement to know Ψ(x). A single case picks out only one component
< x | Ψ >, actually one experiment tells us only that Ψ(x0 ) is nonzero, not even relative probability.
CHAPTER 1. HISTORICAL AND EXPERIMENTAL ORIGINS 35

1.16 Derivation of the Schrödinger Equation – Free particles


We can used the superposition principle to form wave packets as initial free states representing our
initial electron. Now we introduce the Einstein – de Broglie relations

E = h̄ω p~ = h̄~k,

to derive the Schrödinger equation. For a nonrelativistic free particle E = p2 /2m = h̄2 k 2 /2m = h̄ω;
therefore, our choice of ω(k) is
h̄ 2
ω(k) = k .
2m
Correspondingly, we have phase and group velocities (see H.W.) vφ = p/2m = vg /2, where ~vg = p~/m
is the group or particle velocity. Now we can see that
~ ~
e+i(k·~r−ω(k) t) ~
Z
Ψ(~r, t) = dk φ(~k − ~k0 )
(2π)3/2
satisfies
~ ~
h̄2 2 e+i(k·~r−ω(k) t) ~ h̄2 2
Z

(− ∇ − ih̄ )Ψ(~r, t) = dk ( k − h̄ω)φ(~k − ~k0 ) = 0.
2m ∂t (2π)3/2 2m
x
or
H0 Ψ = ih̄Ψ̇,
where
h̄2 2
H0 = − ∇ .
2m
Also,
h̄2 2 ~ ~
k φ(k − k0 ) = h̄ω φ(~k − ~k0 ),
2m
which is our first case of the Schrödinger equation in momentum space. Here we see that p~ is
represented by an operator
h̄ ~
p~ → ∇.
i

1.16.1 The first derivative and why “i” in the Schrödinger Equation
The free particle Schrödinger equation is

∂ h̄2 2
ih̄ Ψ(~r, t) = − ∇ Ψ(~r, t) . (1.33)
∂t 2m
We saw that the

ih̄
∂t
appears as a consequence of the superposition principle and of vφ = 12 vg (or ω = 2m h̄ 2
k . Another
2 ∂
way of saying this is that Schrödinger made a diffusion equation ∇ y(~r, t) = a ∂t yy(~r, t) (diffusion
occurs for real a) into a propagating and spreading equation for Ψ by taking a as imaginary. Another
feature here is that the Schrödinger involves only the first time derivative and thus represents an
initial value problem (IVP), as needed from the prior discussion of sequential measurement and
the requirements of the uncertainty principle. Note that the usual wave equation has both second
2
temporal and second spatial derivatives, ∇2 y = v12 ∂∂2 t y and thus one needs to specify both the
initial function and the initial derivative for the usual wave equation.

Another reason we need i ∂t is to conserve probability flux, in contrast to pure diffusion where
the function vanishes or attenuates.

Thus we note that the ih̄ ∂t structure produces an equation which has both propagation ( as
required by trajectories) and spreading ( as required by the uncertainty principle).
x Note at this stage φ(~k − ~k0 ) is assumed to be independent of time.
CHAPTER 1. HISTORICAL AND EXPERIMENTAL ORIGINS 36


The occurrence of i = −1 in quantum mechanics forms an essential use of complex numbers
and complex wave functions in quantum mechanics, as we shall see repeatedly. Later we shall see
how complex conjugation will be part of our study of time reversal.
Finally another role of “i” is to provide conservation of probability and proper normalization by
the steps:

∂ h̄2 2
Ψ(~r, t) = −
ih̄ ∇ Ψ(~r, t) , (1.34)
∂t 2m
after taking the complex conjugation we also have

∂ ∗ h̄2 2 ∗
−ih̄ Ψ (~r, t) = − ∇ Ψ (~r, t). (1.35)
∂t 2m
Now multiply the first of these by Ψ∗ , the second by Ψ and subtract to deduce

∂ h̄2 ∗ 2
~ h̄ [Ψ∗ (~r, t)∇
~ Ψ(~r, t)−Ψ(~r, t)∇
~ Ψ∗ (~r, t)].
ih̄ (Ψ∗ (~r, t)Ψ(~r, t)) = − [Ψ (~r, t)∇2 Ψ(~r, t)−Ψ(~r, t)∇2 Ψ∗ (~r, t)] ≡ ∇−
∂t 2m 2m
(1.36)
This can be cast into the form of a continuity equation

~ · ~j(~r, t) + ∂ ρ(~r, t) = 0,
∇ (1.37)
∂t
where we identify the probability density by

ρ(~r, t) ≡ Ψ∗ (~r, t)Ψ(~r, t)) ≥ 0, (1.38)

and the probability flux as

~j(~r, t) h̄ [Ψ∗ (~r, t)∇


~ Ψ(~r, t) − Ψ(~r, t)∇
~ Ψ∗ (~r, t)]. (1.39)
2im

If ~j(~r, t) vanishes at an infinite sized sphere (nothing escapes), we can integrate over all space and
deduce that the probability remains normalized to one
Z
d3 rΨ∗ (~r, t)Ψ(~r, t) = 1,

at all times since Z



ρ(~r, t) = 0.
∂t
Note that the Schrödinger equation is really two coupled equations involving the real and imag-
inary parts of Ψ,
∂ h̄2 2
−h̄ ΨI (~r, t) = − ∇ ΨR (~r, t) , (1.40)
∂t 2m
and
∂ h̄2 2
h̄ ΨR (~r, t) = − ∇ ΨI (~r, t). (1.41)
∂t 2m
Chapter 2

Probability and Wave Packets

Last time we learned that the uncertainty principle applies to matter (electrons) and light and re-
stricts our ability to measure x and p simultaneously. We saw how the double slit illustrates that
the particle and wave aspects are complementary; i.e., determination of the trajectory wipes out
the possibility of interference. Interference occurs only if the trajectory (or hole used) remains un-
known. Thus we turned to a statistical or probabilistic approach in which we attribute a probability
to each election, which is determined by repeating the experiment. It was also emphasized that
Einstein’s objections helped to clarify many points in quantum mechanics, but that controversy
remains. Fortunately, most of the controversy was in the philosophic domain, without alteration of
the predictions and use of the formalism.
We will stick to the Copenhagen interpretation which states it is impossible to violate the un-
certainty principle in preparing states, and that the probability description is necessary.
Let us give a rough rendition of the basic postulates of quantum mechanics concerning measure-
ment - I will give a better version later after some added ideas are developed. Then we’ll turn to the
description of initial states prepared to be consistent with the demands of the uncertainty principle.

2.1 Statistical Interpretation


Let us note the following basic assumptions of quantum mechanics. The postulates will be expressed
more completely later.
We now describe the system by a wave function Ψ, which is a probability amplitude. All infor-
mation about the system is stored in Ψ and means of handling that information is the job of the
theory. Several points should be noted:
1. Ψ represents the probability amplitude, the probability of finding the electron(s) at some
point(s) is
P = Ψ∗ Ψ ≥ 0 - with d3 r P = 1 The above definition yields interference.
R

2. Ψ refers to a single electron - to measure it many repeated experiments are needed.


3. Ψ is an abstract wave in configuration space; for example, the N-particle system Ψ(r, ...rN , t)
is described in the 3N-dimensional space ≡ phase space or configuration space - not to be
confused with an actual material wave in space. Use of a universal t is a non-relativistic
feature.
4. Ψ obeys a superposition principle Ψ = Ψ1 + Ψ2 , where Ψ1 and Ψ2 represent two possible states
of the N-electron system. The above linear combination is another possible state - leads to
interference.
5. Eigenvalues are always observed in an individual measurement (see later).
Now there are historically two ways to proceed:
1.) 1925 Heisenberg - use of Rydberg-Ritz plus the Correspondence Principle led Heisenberg
along with M.Born and P. Jordan, to Matrix Mechanics. His viewpoint was to deal with observables
only, such as the frequency and amplitude of radiation. Close to what we do today.
CHAPTER 2. PROBABILITY AND WAVE PACKETS 38

2.) 1926 Schrödinger’s Wave Mechanics - based on analogy with waves/optics and the Hamilto-
nian dynamics.

2.2 Wave Packets


2.2.1 General Remarks
To do experiments with atomic systems, we must first prepare the initial conditions (say the electron
or molecular beam) using slits, an accelerator, etc. The preparation of the initial state is, however,
subject to the limitations imposed by the uncertainty principle. The initial position and momentum
can not be simultaneously determined with arbitrary precision and we wish to express that limitation
on our ability to prepare an initial state in a convenient mathematical language.
It should be stressed that it is of practical significance that we remove from our formalism any
artificial claims about knowing ∆px and ∆x to be small simultaneously, just as it is of practical
significance in special relativity to free the formalism from artificial notions of absolute space, aether,
or universal time. The practical consequence of our expressing our limited knowledge of our initial
state is that a wide range of phenomena can be understood and predicted by such a formalism. That
is, the theoretical framework is improved in its job of organizing our means of relating one set of
facts to another and in having predictive power. You see honesty pays, at least in theory.
The description of our initial state, which we consider to be a beam of free, noninteracting
electrons is expressed in terms of the wave function of each electron Ψ(x), which is the probability
amplitude used to construct the probability of finding an electron between x and x + dx.

P (x) = Ψ∗ (x)Ψ(x).
Of course, to measure Ψ(x) many repeated observation are needed, despite the fact that Ψ(x) is an
attribute of each electron. Recall also that Ψ is an abstract mathematical probability wave - not a
material wave in real space.
The construction of wave packets is based on the superposition principle, which is a basic postu-
late of the quantum theory needed to generate the observed interference and diffraction properties
of electrons. The wave packets are built from waves but come close to the particle concept by having
the possibility of being localized to the extent permitted by the uncertainty principle. The basic
problem of understanding observed trajectories is confronted here. a
After preparation of the initial state of a free electron, the typical experiment involves a further
interaction of the system with a target (or in the case of a bound state a continuing interaction).
That requires knowing the dynamics of how the probability evolves in time; i.e., it will lead us to
full the Schrödinger equation. For now, let us restrict our discussion to the single free electron. How
can we describe it?

2.2.2 Mathematical Construction of Wave Packets


b
The superposition principle permits us to construct various types of localized states using the
general Fourier transform in three-dimensions
~ ~
e+i(k·~r−ω(k) t) ~
Z
Ψ(~r, t) = dk φ(~k − ~k0 )
(2π)3/2
applied to a single electron case c . The angular frequency ω(k) is in general a scalar function of
~k. For simplicity we will take ω(| ~k |) to be a real scalar function of | ~k |, which thus represents
a i.e.
of the paths seen in Wilson cloud chamber.
b We could introduce other variables in φα , such as the spin, isospin, · · · other quantum numbers to characterize
the state. Also, other “hidden parameters” could be used to parameterize φα to make each wave packet distinctive
and to attempt to eliminate the stochastic aspect of events. Various types here also refers to various wave packet
shapes.
c Which requires that Ψ be normalized as:
Z Z
Ψ∗ (~
r, t) Ψ(~
r, t) d~
r=1= d~k φ∗ (~k − ~k0 ) φ(~k − ~k0 )
CHAPTER 2. PROBABILITY AND WAVE PACKETS 39

plane wave motion in a dispersive, nondissipative medium. The quantity φ weights the various wave
lengths λ– = k −1 in a manner that generates various types of localized wave packets Ψ. (Later we
will refer to e−iωt φ as the wave function in the momentum representation). We can invert (1) for
t = 0 and find
~
e−ik·~r
Z
φ(~k − ~k0 ) = d~r Ψk0 (~r, 0),
(2π)3/2
where we have used a standard Fourier transform result which is conveniently expressed in terms of
the 3-D delta function (Dirac)d
~0 ~
ei(k −k)·~r
Z
δ(~k 0 − ~k) = d~r = δ(kx0 − kx )δ(ky0 − ky ) δ(kz0 − kz ).
(2π)3
It is also possible to form wave packets for two free electrons, say prepared in a colliding beam
situation, using
Z
1 ~ ~
Ψ(~r1 , ~r2 , t) = ei(k1 ·~r1 −ω1 t) ei(k2 ·~r2 −ω2 t) φ1 (~k1 − ~k01 ) φ2 (~k2 − ~k02 ) d~k1 d~k2 .
(2π)3
Note the above case assumes a nonrelativistic situation (common time). Generalization to N > 2
particles is obvious (say to d, d scattering). Also one needs to antisymmetrize properly for colliding
Fermions.
A special case of interest is the plane wave (representing an electron of definite momentum)
where
~
k0 ·~
r −ω(k0 ) t ]
Ψ(~r, t) = ei[ /(2π)3/2
and
φ(~k − ~k0 ) = δ(~k − ~k0 )
Note that for this case ∆k → 0, but ∆x → ∞. Momentum is conserved so later development (t >
0) is clearly as given above.
Another important case is the situation where an electron is initially located at ~r = ~r0 , for whiche

Ψ(~r, 0) = δ(~r − ~r0 )

and
~
φ(k − k0 ) = e−ik·~r0 /(2π)3/2
Note that now ∆k → ∞ and ∆r → 0. At later times the wave function is
Z i[~k·(~r−~r0 )−ω(k)t]
e
Ψ(~r, t) = d~k
(2π)3
At t > 0 the presence of e−iω t shows that Ψ(~r, t) is no longer localized at ~r0 but in general spreads.
For the special cases of:
(1) ω(k) = a constant, independent of k
and
(2) ω(k) = ~k · ~vφ (which applies to light) there is no spreading. Instead, we get for (2)
Z
1 ~
Ψ(~r, t) = eik·(~r−~r0 −~vφ t) d~k = δ(~r − ~r0 − ~vφ t),
(2π)3
which represents just a move to ~vφ t + ~r0 with no change in shape. Case (1) yields Ψ(~r, t) =
e−iωt δ(~r − ~r0 ) –no motion, just a phase change –an unusual case! We see that spreading occurs only
for a dispersive medium, which means ω depends on k in a way different from (2) or (1).
We can now describe the general situation corresponding to any initial wave packet, not just the
two extreme cases considered above. We have
~
eik·~r
Z
~
Ψ(~r, 0) = 3/2
φ(~k − ~k0 ) d~k = eik0 ·~r A(~r),
(2π)
d Properties of the Dirac delta function are given separately.
e This is not a properly normalized state; think of it as a sharply peaked localized function.
CHAPTER 2. PROBABILITY AND WAVE PACKETS 40

~
eik·~r
Z
A(~r) = modulation factor = φ(~k) d~k
(2π)3/2
or
~
e−ik·~r 3
Z
φ(~k) = d r A(r)
(2π)3/2
Hence the wave packet develops as f
Z i(~k·~r−ω(k) t) ~ ~ 0
e ~ e−i(k−k0 )·~r
Ψ(~r, t) = dk d~r 0 A(~r 0 ),
(2π)3/2 (2π)3/2
Z i(~k· (~r−~r 0 )−ω(k) t )
e
= d~k Ψ(~r 0 , t0 ) d~r 0 ,
(2π)3
Z
Ψ(~r, t) = (~r | G0 (t − t0 ) | ~r 0 ) Ψ(~r 0 , t0 ) d~r 0 ,
where the Green’s function for this general Huyghen’s construction procedure is
~ −i~ r 0
eik·~r k·~
Z
−iω(k)(t−t0 ) e
0
(~r | G0 (t − t0 ) | ~r ) = e d~k (2.1)
(2π)3/2 (2π)3/2
i
= < ~r | e− h̄ H0 (t−t0 ) | ~r 0 ) −→ δ(~r − ~r 0 ).
t→t0
0 0
Note for t = t0 , Ψ(~r , 0) = δ(~r − ~r0 ), the previous case is recovered with
Ψ(~r t) = (~r | G0 (t − t0 ) | ~r0 ).
Later this will all be expressed formally as
H0
| Ψ(t) >= e−i h̄ (t−t0 )
| Ψ(t0 ) >= G0 (t − t0 ) | Ψ(t0 ) >

ih̄ | Ψ >= H0 | Ψ >
∂t
i
G0 (t − t0 ) = e− h̄ H0 (t−t0 )
0
< ~r | G0 (0) | ~r >= δ(~r − ~r 0 ) ,
all of which comes after we specify H0 – so far ω(k) is general. We will evaluate this Green’s function
for various choices of ω(k) later. Now let us ask some general questions about the development of
the wave packet.
Our general wave packet can also be written as
Z i(~k·~r−ω(k) t)
e
Ψ(~r, t) = = d~k φ(~k − ~k0 ) (2.2)
(2π)3/2
Z i[ (~k+~k0 )·~r−ω(~k+~k0 ) t ]
e
= φ(~k) d~k.
(2π)3/2
Also,
~ ~
Ψ(~r, t) = ei (k0 ·~r−ω(k0 )t) A(~r, t)
Here
~
ei[ k·~r−δω t ] ~
Z
A(~r, t) = φ(k),
(2π)3/2
with δω ≡ ω(~k + ~k0 ) − ω(~k0 ).
Note again that for ω(k) = ~vφ · ~k, the nondispersive case, we get
Z i~k·(~r−~vφ t)
e
A(~r, t) = φ(~k) d~k (2.3)
(2π)3/2
= A(~r − ~vφ t, 0) (2.4)
= A(~r − ~vφ t) .
There is no change in shape and the shape A is moving with a velocity given by vφ . Again this
applies to light in vacuum. For this case ~vφ = ~vg .
f We set t0 ≡ 0.
CHAPTER 2. PROBABILITY AND WAVE PACKETS 41

2.2.3 Some Features of Wave Packets


Before we proceed to discuss the time development of the wave packet, let us picture it; the packet
is composed of a plane wave and a modulation.
~
Ψ(~r, t) = ei(k0 ·~r−ω0 t) A(~r, t).

Normalization requires that Ψ∗ Ψ dr = 1 or A∗ (~r, t)A(~r, t) d3 r = 1. For the case of k0−1 =


R R
–0 << width of A, we have at t = 0, ∆r > λ
λ –0 , where ∆r is the size of the packet. Thus, A serves
to form a bundle, envelope, or packet in which the plane wave oscillations are restricted.
Associated with ∆r, the function φ(k) has a width ∆k, which is ∆k ∼ 1/∆r or ∆p ∆x ∼ h̄.
Thus we are storing the uncertainty principle in a useful way. How fortunate that Fourier transforms
have just that property!
It is worth noting that the 3-D wave packets can have various shapes. If we pick φ(~k − ~k0 ) to be
spherically symmetric, then the initial wave packet will really be a ball, a wave packet ball, which
is as close to being localized as we can get. If ω is taken as ω(| ~k |), which is the usual choice,
then spherically symmetry will be preserved at later times, although the size of the sphere will
increase.
q If ω(kz ) is used, then a distortion of the shape will occur. Also for a relativistic packet
(ω = k 2 c2 + m2 c4 /h̄2 ) a distortion occurs for t > 0 - see later. Of course, it is possible to produce
nonspherical initial wave packets.

2.2.4 One Dimensional Wave Packets


Instead of dealing with spherical wave packets, it is often convenient to study one-dimensional wave
packets. In analogy to the three dimensional case, the case of motion in the z− direction, we have
Z i(kz z−ωz t)
e
Ψ(z, t) = √ φ(kz − k0z ) dkz , (2.5)

from which it follows that
e−ikz z
Z
φ(kz − k0z ) = √ Ψ(z, 0) dz, (2.6)

and as before Z
Ψ(z, t) = (z | G(t − t0 ) | z 0 ) Ψ(z 0 , t0 ) dz 0 . (2.7)

Now the 1-D Green’s function is


0
e−ikz z iωz (t−t0 ) eikz z
Z
0
(z | G(t − t0 ) | z ) = √ e √ dkz . (2.8)
2π 2π
The 1-D case can also be obtained by simply considering the particle localized in the x − y plane
and setting
φ(~k − ~k0 ) → φ(kz − k0z ) δ(kx ) δ(ky ), (2.9)
taking k0x = k0y = 0, for convenience.
The question now is: What happens later (t > 0) after preparing these packets. One way of
answering this question is to introduce the idea of group velocity.

2.2.5 Group Velocity


The idea of using a group velocity to describe the motion of the wave packet (or of A(~r, t)) is an
approximation based on expanding δω = ω(k + k0 ) − ω(k0 ) as


δω = ω(k + k0 ) − ω(k0 ) = 0 + k ( ω(k)) |k0 + · · · ,
∂k
or for the 3–D case

~ k ω(k)] |k + ki kj [∇k ∇k ω(k) |k ] + · · · .


δω = ω(~k + ~k0 ) − ω(~k0 ) = 0 + ~k · [∇ 0 i j 0
2
CHAPTER 2. PROBABILITY AND WAVE PACKETS 42

~ k ω(k) |k and also define ρij ≡ ∇k ∇k ω(k) |k and have


Now we define the group velocity ~vg = ∇ 0 i j 0

ki kj
δω = ~k · vg + ρij + · · · .
2
To see how ~vg describes the motion of the packet use this expansion as follows
~
Ψ(~r, t) ei(k0 ·~r−ω0 t) A(~r, t)
=
Z i(~k·~r−δω t)
e
A(~r, t) = φ(~k) d~k
(2π)3/2
Z
1 ~ ~ ki kj
= 3/2
ei(k·~r−k·~vg t) e−i 2 ρij t φ(~k) d~k
(2π)
Z i~k·(~r−~vg t) Z i~k·(~r−~vg t)
e ~ 1 e
≈ 3/2
φ(k) dk + ρij t ki kj φ(~k)d~k,
(2π) 2i (2π)3/2
i
= A(~r − ~vg t; 0) + ρij t ∇ri ∇rj A(~r − ~vg t; 0) + · · ·
2
1
= [1 − t ρij ∇ri ∇rj ]A(~r − ~vg t; 0) + · · ·
2i
Note that there are two expansions being used above, one is the expansion of δω, the other is the
expansion of the exponential in a small-spreading approximation. The first expansion is often not
even needed, the second is used to establish criteria for spreading. Note the leading term has the
same shape as at t=0, the packet moves to ~r → ~r + ~vg t. Thus vg describes motion of the packet,
provided higher order terms are small. The normalization is also maintained, but the exponential
expansion needs to be avoided to maintain proper normalization.

Notes on Group Velocity and Index of Refraction


1.) For Light
c c
vφ = √ =
µ n

n= µ = n(k) = n(ω) = index of refraction

kc nω
ω = kvφ =
; k=
n c
vφ < c if n > 1 usual case, but vφ > c if n < 1 can happen.
Group velocity
∂ω ∂k n ω ∂n −1
vg = = 1/ =[ + ] .
∂k ∂ω c c ∂ω
dn
For anomalous dispersions dω < 0 and large.

ω ∂n −1
vg = vφ [1 + ]
n ∂ω
Hence, we could get vg > vφ and vg > c - which indicates a breakdown of the group velocity idea;
i.e. of the series expansion for δω in which case turn to ideas of signal/forerunner and precursor
velocities, see L. Brilliouin “Wave Propagation and Group Velocity.”
Also note that:
kc
ω = kvφ =
n
∂vφ – ∂ vφ
... vg = vφ + k = vφ − λ –
∂k ∂λ
∂vφ
vg > vφ if >0
∂k
CHAPTER 2. PROBABILITY AND WAVE PACKETS 43

∂vφ
vg < vφ if <0
∂k
2.) Q. M. Case (Non-Relativistic)
ω h̄k 1 c
vφ = = = vg = ;
k 2m 2 n
∂ω p
vg = =
∂k m
∂vφ h̄
= >0
∂k 2m
... vg > vφ ; vg = 2vφ
2mc 2
n= = – ,
h̄k λe k
dn n dk
=− <0
dω k ∂ω

... vg = or vg = 2vφ
1 − vφ /vg

2.2.6 Estimate of Spreading


We have found that

~ 1
Ψ(~r, t) ≈ ei(k0 ·~r−ω0 t) [1 − t ρij ∇ri ∇rj ] A(~r − ~vg t, 0)
2i
+ · · · higher order terms involving ∇nk ω(k)
the initial shape motion is determined by the quantities ~vg = ∇ ~ k ω(k) |k and ρij = ∇k ∇k ω(k) |k .
0 i j 0
~ n
For ω(k) = k · ~vφ , we have ρij = 0 and (∇ · · · ∇) ω = ρij···n = 0! Thus, no change in shape occurs
for ω(k) = ~k · ~vφ for which vg = vφ , which applies to light in vacuum.
h̄ 2
For the choice ω = 2m k , we find
p~ ~vg
~vφ = = ;
2m 2
p~
~vg = = ~vparticle ,
m

and vg vφ = 12 vg2 and ρij = m δij .
g
The relativistic case causes the shape of a rapidly moving spherical wave packet to take on a
g For the relativistic case: p
E = h̄ω = p 2 c2 + m 2 c4
and hence p
ω= k2 c2 + kc2 c2 .
There follows: ω p
vφ = = c 1 + kc2 /k2
k
1/2
vg = p 2kc2
k c2 + kc2 c2
2

mc –−1
kc = =λ c

kc2 c 2
vg = = ; vφ vg = c2 .
ω vφ
1 k j c2 c2 1 2c2 ki
ρij = ∇i ∇j ω = ∇i 2kj c2 = ∇i = δij + kj c2 (− ) 3
2ω ω ω 2 ω
c2 c4 1
ρrel
ij = δij − ki kj 2
ω ω ω
c2
= vg /k
ω
ki kj
ρ = δij vg /k − (vg /k)2
ω
CHAPTER 2. PROBABILITY AND WAVE PACKETS 44

“pancake” shape due to relativistic contraction. h .


For simplicity, let us take the initial packet as
1 2 2 2 2
A= e−r /2a = N e−r /2a
(πa2 )3/4

With a = 2∆r
1 2 2
A2 = [ 2 ]3/2 e−r /a
πa
r
1 2 −r2 /2a2
lim 3 e = δ(r)/r2 .
a→0 a π
We have found for the nonrelativistic choice of ω that
~ 1 h̄
Ψ(~r, t) ' ei(k0 ·~r−ω0 t) [ 1 − t ∇2 ] A,
2i m
for a Gaussian shape

~ 2 2 2 3 1 1
∇2 A = N ∇(− 2
~r e−r /2a ) = (− 2 + 2 ~r 2 2 )A
2a a a a
1 r2
= [ − 3]A
a2 a2
~ i h̄ t r − vg t 2
... Ψ(~r, t) = ei(k0 ·~r−ω0 t) {1 + 2
[( ) − 3 ] }A(~r − ~vg t), 0).
2ma a
The degree of spreading is determined by the dimensionless parameter
t h̄ 2 1
e= ( ) . (2.10)
h̄ a 2m
With that parameter and defining ~r(t) = ~r − ~vg t, we have

~ ~r(t) 2
Ψ = ei(k0 ·~r−ω0 t) {1 + ie[( ) − 3]} A(~r(t), 0).
a
The associated probability is

~r(t) 2
P (~r, t) = A2 (~r(t), 0)[1 + e2 [( ) − 3]2 ] + · · · .
a
At the points r(t) = ±a, the probability is enhanced by the factor [1 + e2 (4)], which indicates the
increased size of the wave packet.
~r − ~vg t 2
Ψ∗ Ψ = P (~r, t) ' A2 (~r − ~vg t, 0) [1 + e2 [( ) − 3]2 ] + · · ·
a
We require that
t h̄ 2 1
e= ( ) << 1 . (2.11)
h̄ a 2m
to get small spreading!
Here t ∼ time of full experiment ' Lv where L = length of path in experiment. Therefore we can
express this dimensionless parameter as
L h̄ Lλ–B 1 1 L λB
e' = =
a 2mav a a 2 4 ∆x ∆x
vg 2 k ki kj
ρij = ( ) (δij − );
k vg ω
vg
ρi6=j ∼ ( )2 ki kj /ω.
k
Above results are all for a relativistic choice of ω(k) and are also true in the massless limit–light.
h The wave packets built here are still not fully relativistic, since causal light–cone considerations require the

introduction of negative energy states.


CHAPTER 2. PROBABILITY AND WAVE PACKETS 45

L λ –B 1
e'
∆r ∆r 4
a
where ∆r = 2 = size of original packet. If ∆r >> λ
√ –B (or ∆r >> k0 −1 ), which is a typical
situation, perhaps e could be small. Usually L > ∆r so we can still get e ≈ small.
Note that this result can also be obtained directly from the uncertainty principle using the steps
(see Homework)
∆px h̄/2
∆x(t) ∼ ∆x(0) + t ∼ ∆x(0) + t,
m ∆x(0)m
or

∆x(t) ∼ ∆x(0)[1 + t] = ∆x(0)[1 + 2],
2m∆x(0)2

where we identify  = 4m∆x(0)2 t, as before.

Spreading in Orbits
If we try to localize an electron in orbit, its wave packet spreads rapidly for lower levels. For higher
n-values the spreading is less. This is the correspondence principle, now expressed in terms of wave
packet properties. To see this result consider an electron localized to a small region, so that its ∆r is
very small compared to the radius of its orbit, rn . For example, take ∆r = rn /10, where the orbital
2
radius for a Bohr atom is rn = nZ a0 in terms of the Bohr radius a0 . How much will the localized
electron wave packet spread as it makes one revolution over the length L = 2πrn ? We also have
–B = h̄ = rn . Therefore, the measure of the spreading is determined by
λ mv n
rn
L λB 2πrn n
e∼ ∼ rn × rn ,
∆r ∆r 10 10

which reduces to
10 200π
e ∼ 20π∼ .
n n
This estimate of spreading shows that for low n orbits, the wave packet already spreads out over
the whole atom in one revolution ! For very large n the spreading can become small and one then
has the classical, correspondence principle limit.

2.2.7 Review
We discussed several aspects of wave packets; namely, the notion of group velocity, and an estimate
of spreading. We found that:

~ 1
Ψ = ei(k0 ·~r−ω0 t) [1 −t ρij ∇i ∇j ]A(~r − ~vg t, 0) + · · ·
2i

~
~vg ≡ ∇ω(k) |k = ~k = ~vparticle for NR
m

ρij = ∇ki ∇kj ω(k) |k0 = δij for NR,
m
for the nonrelativistic (NR) case.
We found that an initial Gaussian packet evolves in time as
(~r − ~vg t)2
P (~r, t) =| Ψ |2 = A2 (~r − ~vg t, 0)[1 + e2 [ − 3]2 ],
a2
which is the exact NR result. This confirms our result that the condition to get small spreading is:
t h̄ 1 1 L λ–B
e = ( )2 ∝ << 1.
h̄ a 2m 4 ∆r ∆r
Now let us discuss the Gaussian wave packet some more; we will be able to obtain a better form
for that case and see the explicit spreading effect. We will then go on to derive the Schrödinger
equation for a free particle.
First here are some useful integrals, followed by a derivation of the above result. The one-
dimensional case is in the Homework.
CHAPTER 2. PROBABILITY AND WAVE PACKETS 46

Useful Integral
Z ∞ Z ∞
1 B2
−(Ak2 +Bk) 1 B 2
e dk = dk e−A(k+ 2 A ) e+ 4 A

−∞ −∞

B 1 B 2 1 B2
Ak 2 + Bk = A(k 2 + k) = A{(k + ) − }.
A 2A 4 A2
Z ∞ r
1 B2 2 1 B2 π
= e4 A dk e−A k = e 4 A
−∞ A

A=i t; B = −ix
2m
ih̄
Ak 2 = iω(k)t ⇒ A = t
2m

Z r
1 x2 π
ei(kx−iω(k)t) dk = e− 4 ih̄t 2m 2m
−∞ ih̄t
−∞
ei(kx−ω(k)t)
Z
ix2 m m 1/2
... dk = e+ 2h̄t ( )
−∞ (2π) 2πih̄t
For the 3–D case:
~
ei[k·~r−ω(k)t]
Z
m 3/2 imr2
d~k 3
=( ) e 2h̄t
(2π) 2πih̄t
with
h̄ 2
ω= (k + ky2 + kz2 )
2m x
Also: Z
2 ~ r) ~2
1 B π
d~r e−(Ar +B·~
= e4 A ( )3/2
A

2.3 Gaussian Wave Packet


Z
i~
k0 ·~
r
Ψ(~r, 0) = e A(~r, 0) with | A |2 d~r = 1

1 2 2
A(~r, 0) = e−r /2a
(πa2 )3/4
2
a ~ ~ 2a 2
φ(~k − ~k0 ) = ( )3/4 e−(k−k0 ) 2
π
~
Ψ(~r, t) = ei(k0 ·~r−ω0 t) A(~r, t)
Direct Method
~ ~
ei[(k−k0 )·~r−(ω−ω0 )t] ~ ~
Z
A(~r, t) = φ(k − k0 ) d~k (2.12)
(2π)3/2
~ h̄ 2 ~ ~
ei[k·~r− 2m (k +2k0 ·k)t] a2 3/4 −k2 a2 /2
Z
= ( ) e (2.13)
(2π)3/2 π
Z i[~k·R−ω(k)T
~
a2 3/4 e ]
= ( ) dk (2.14)
π (2π)3/2
m 3/4 i mR2 a2 3/4
= (2π)3/2 ( ) e 2h̄T ( ) (2.15)
2πih̄T π
h̄ ~
R = ~r − k0 t = ~r − ~vg t
m
ih̄ h̄t
T = i + a2
m m
CHAPTER 2. PROBABILITY AND WAVE PACKETS 47

1 2m a2
T =t+
i h̄ 2
1 ih̄t −3/2 −(~r−~vg t)2 /2(a2 + ih̄ t)
A(~r, t) = [ a+ ] e m
π 3/4 ma
1 2 2
A(~r, 0) = 2 3/4
e−r /2a
(πa )
1 2
/ā2
P (~r, t) = Ψ∗ Ψ = e−(~r−~vg t) ;
(πā2 )3/2
h̄t 2
with ā2 ≡ a2 + ( ma ) ; ā2 = a2 + 4a2 e2 = a2 (1 + 4e2 ).

2.3.1 Greens Function - Gaussian Wave Packet


Z
Ψ(~r, t) = G(~r, t | ~r 0 , t0 ) d~r 0 Ψ(~r 0 , t0 )
Z
m 3/2 1 i~ r2
r − 2a
k0 ·~ m 02 ~ ·~ 0 1 02
r 0)
= ( ) e e 2
ei 2h̄t r eik0 r
e− 2a (r r ·~
+2~
d~r 0 .
2πih̄t (πa2 )3/4
Z
02 ~ r 0) 1 B2 π 3/2
e−(Ar +B·~
d~r 0 = e 4 A (A )

im 1
A=− + 2
2h̄t 2a
~ = −i~k0 + 1 ~r
B
a2
m 3/2 π 3/2 ~ r2 1 1 r 2 1 im
Ψ(~r, t) = ( ) 2 3/4
eik0 ·~r e− 2a2 1 im 3/2
e 4 ( a2 −ik0 ) /( a2 − 2h̄t )
2πih̄t (πa ) [ 2a2 − 2h̄t ]
r −~ 2
ih̄t −3/2 i(~k0 ·~r−ω0 t) − 2(a(~ vg t)

Ψ(~r, t) = π −3/4 (a + ) e e 2 +i(h̄/m)t)

ma
as before.

2.3.2 Gaussian Packet: (continued)


Later we shall define the fluctuations or root mean square deviations using:
Z
(∆rx ) = P (~r, t) d~r [rx − < rx >]2 =< rx2 > − < rx >2
2

and Z
(∆px )2 = h̄2 P (~k, t) d~k [kx − < kx >]2 = h̄2 [< kx 2 > − < kx >2 ]

where P (~k, t) =| φ(~k − ~k0 ) |2 is independent of time and is the wave function in momentum space.
Evaluate Integrals
Z ∞ Z ∞ −(rx −vgx t)2 /ā2
1 2 2 e
< rx >= rx d~r 2 3/2
e−(~r−~vg t) /ā = (vgx t + rx0 )
−∞ (πā ) ∞ (πā2 )1/2
h̄t 2 h̄t
ā2 = a2 + (
) = a2 (1 + 4e2 ); e =
ma 2ma2
We obtain the classical result < rx >= vgx t + 0 and similarly < px >= h̄k0x . Now consider:
2 2
e−(rx −vgx t) /ā
Z Z
2
< rx 2 >= rx 2 drx = vgx 2 t2 + drx rx 2 e−(rx /ā) /(πā2 )1/2
(πā2 )1/2

ā2
< rx2 >= vg2 t2 +
2
CHAPTER 2. PROBABILITY AND WAVE PACKETS 48

ā2 (t)
or < rx2 (t) >= vg2 t2 + 2 . Now for the momentum part:

a2 a2
Z Z
2 2 2
2
< px >= px 2
dkx ( )1/2 e−(kx −k0x ) /a = (h̄k0x )2 + h̄2 ( )1/2 dkx kx 2 e−(kx a)
π π

h̄2
= (h̄k0x )2 +
2a2
Note the original a appears here, not ā which depends on time.
Combining the two results, we have (∆rx (t))2 = ā2 /2, (∆px (t))2 = h̄2 /(2a2 and hence the
product
h̄ ā h̄ h̄
∆rx (t) ∆px = = [ 1 + 4e2 ]1/2 ≥ .
2 a 2 2
At t = 0 this becomes ∆rx (0) ∆px = h̄2 , which is the best we can do to satisfy the uncertainty prin-
ciple. The Gaussian packet is therefore called the minimal packet because it satisfies the uncertainty
principle at least at t = 0 with the equal sign. Thereafter the spreading causes the inequality to be
applicable.
Chapter 3

Schrödinger Equation–Free
Particle Motion

In the prior sections, we studied the spreading of a Gaussian wave packet using the Green’s function
and found
1 2 2
| Ψ(~r, t) |2 = 2 3/2
e−(~r−~vg t) /ā
(πā )
with ā2 = a2 (1 + 4e2 ). To assure small spreading, we required that

h̄t
e= << 1.
2ma2
We also obtained
h̄ ā h̄
∆rx (t) ∆px = = [1 + 4e2 ]1/2 ≥ h̄/2
2a 2
√ h̄
∆rx = ā/ 2 ∆px = √ .
2a
Then we derived the Schrödinger equation for a free particle and found that

p2 ∂
Ψ(~r, t) = ih̄ Ψ(~r, t)
2m ∂t
with the association
h̄ ~
p~ → ∇.
i
The Schrödinger equation follows from the: a) Superposition principle, and from b) The Einstein –
de Broglie relations E = h̄ω = p2 /2m and p~ = h̄~k or vg = 2vφ .
∂ ∂
Then we discussed the questions: Why i? Why ∂t ? The answer is first of all i ∂t appears because

of a) and b) above. Secondly, we turn a diffusion equation into a wave equation by taking i ∂t . The
00 00
diffusion equation: y = −y becomes complex y = −i ẏ ; and a complex diffusion equation has
the needed property of both propagating and spreading ! The first–order time derivative assures us
that only Ψ(~r, 0) is needed to specify later states and not the initial time derivative Ψ̇(~r, 0), as is
required for the usual wave equation, y 00 = − 1c ÿ. That is we have an initial value problem. Finally,

the i ∂t operator is needed to give us a proper continuity equation for the flux of probability.
Today, we will discuss the continuity equation - apply it to the Gaussian packet and then talk
about expectation values and Hermitian operators.

3.1 Probability Flux



The −1 appearing in the Schrodinger equation causes the wave function to be in general complex
Ψ∗ 6= Ψ. There are therefore two equations

ih̄Ψ̇ = H0 Ψ
CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 50

−ih̄Ψ̇∗ = H0 Ψ∗
p2 h̄2 ~ Complex conjugation means i → −i
where H0 is real, H0 = 2m → − 2m ∇2 using the rule p~ = h̄i ∇.

everywhere it appears (H0 = H0 ).

As a result of these equations, with the i ∂t factor playing a crucial role (see H.W. on Klein-
∂2
Gordon equation for the situation with ∂t2 ), we have

∂ ∗ h̄2 ~ ∗ ~ ~ ∗]
ih̄ Ψ Ψ = Ψ∗ H0 Ψ − ΨH0 Ψ∗ = − ∇[Ψ ∇Ψ − Ψ∇Ψ
∂t 2m
or identifying ρ = Ψ∗ Ψ ≥ 0 as the probability density and

~j = h̄ (Ψ∗ ∇Ψ ~ ∗ ) = h̄ Im(Ψ∗ ∇Ψ).


~ − Ψ∇Ψ ~
2im m
a
as the probability current flux. Note if Ψ is real no current occurs (which is OK for bound states).
We need to have a current for the description of scattering or continuum problems and therefore Ψ

must be complex and i ∂t is needed.
Let us apply this expression to a few simple cases.

3.1.1 Electron-charge
We can associate the above probability and probability flux with the charge density and electric
current flux by the simple steps:

−ep ≡ ρ charge densityinCoulombs/Volume

e
− ~j = ~j currentfluxinamperes/area
c
1 ∂ρ ~ ~
... + ∇ · j = 0 = ∂µ j µ
c ∂t
which is current conservation, which is needed for application to electromagnetism.

3.1.2 Plane Wave


A plane wave has the form
~
Ψ = Aei(k·~r−ω t)
[units : Ψ ∝ 1/vol1/2 Φ ∝ vol1/2 ],

with a fixed amplitude A. The probability density is then:

ρ =| A |2 = constant = particles/volume.
The probability current flux is then

~
~j = h̄ | A |2 2 i~k = h̄k | A |2 = ρ ~v = particles / (cm2 − sec)
2im m

3.1.3 Flux for a Wave Packet


A wave packet has the form
~
ei(k·~r−ωt) ~ ~
Z
i(~
k0 ·~
r −ω0 t) ~
Ψ=e A(~r, t) = φ(k − k0 ) dk
(2π)3/2
, where A is a spatial and temporal modification of the amplitude which forms the packet. The
probability density remains simple:
a The current flux can also be written as ~j = Re[Ψ∗ p~op /m Ψ], which as the ρv form.
CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 51

ρ =| A(~r, t) |2 lim ρ → 0,
r→∞

but the current now consists of two terms:


~j = jpacket + jspread ,

where
jpacket = ρ(~r, t) ~v0 .
and
h̄ ~ − (∇A~ ∗ )A)}.
jspread ={A∗ ∇A
2im
Note that since A = A∗ for a Gaussian packet at t=0

~jspread (~r, 0) = h̄ (A∗ ∇A


~ − (∇A
~ ∗ )A) = 0.
2im
For t > 0, A acquires a imaginary part
1 h̄t −3/2 −(~r−~vg t)2 /2(a2 + ih̄ t)
A= (a + i ) e m
π 3/4 ma
and therefore
r̄ 2 r̄ 2
~ = 1 − − −2(~r − ~vg t)
A∗ ∇A e 2(a2 +i h̄t )
m e 2(a2 −i h̄t )
m [ ]
π 3/2 2(a2 + ih̄
m t)
2 2

~ = Im 1 1 r − ~vg t)
− 4 r̄ ah̄t 2 (~ h̄
ImA∗ ∇A 3/2 h̄t
e (a +(
m
) )
2
(−)(a2 − i t)
π 2 2
[a + ( ma ) ]4 a m
h̄ 1 1
= t(~r − ~vg t) 2 [ 2 h̄t 2
] |A(~r − ~vg t)|2
m a a + ( ma )
2
~jspread = h̄ t(~r − ~vg t) 1
h̄t 2 a2
| A |2 → 0 at t = 0
m2 (a2 + ( ma ) ]
h̄ e(~r − ~vg t) h̄t
=2 ρ with e =
m a2 [1 + 4e2 ] 2ma2
h̄ 2 e
j = ρ[~v0 + 2 a ( )(~r − ~vg t)]
m 1 + 4e2
Note that jspread is positive atr > vg t and negative at r < vg t. For small spreading current, e << 1
and the current reduces to the packet current only j = jpacket = ρv0 . The spreading current is
jspreading << jpacket in usual cases. Also note that the effective velocity j = ρ vef f with ~vef f =
2h̄ e
~v0 + ma r − ~vg t) can exceed c for very large r.
2 1+4e2 (~

Thus we see this wave packet has a very high speed front, which would not occur in a proper
relativistic treatment. Fortunately, this effect is negligible in nonrelativistic cases. Since | A |2 = ρ
is very small at large r and also  << 1.. See Gottfried for a discussion of this point. Another
way of seeing that we are limited to nonrelativistic wave packets is by noting that a δ-function or
a wavefunction that vanishes Ψ=0 outside of a sharp edge, e.g. for r ≥ R, yields very high Fourier
components k for which vg → are larger than c. We get currents j corresponding to vef f > c flow;
of course, ρ is very small for such regions but ρ is not analytically exactly zero.

3.2 Probability and Normalization


For large distances the current j → 0. If we assume j falls off faster than the area dS = 4πr2 dr i.e.
j → 0 faster than 1/r2 , we get
Z Z
∂ ~ =0
Ψ∗ Ψ d3 r = − ~j · dS
∂t S→∞
CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 52

and Z
Ψ∗ Ψ d3 r = const = 1

probability is conserved.
Z
Ψ∗ (~r, t) Ψ(~r, t) d3 r = 1

and is normalized if we have all of the system in hand


Z
P (~r, t) d~r = 1

Note: We require both the above fall-off of the current and a Hermitian potential to assure
probability is conserved.

3.3 Expectation Values


The probability P (~r t) = Ψ∗ Ψ can now be used to define average values or expectation values for
the electron’s position. The expectation value of an observable such as ~r is the average value of
numerous measurements and is therefore given by
Z Z
< ~r >= P (~r, t) ~r d~r = Ψ∗ (~r, t) ~r Ψ(~r, t) ≡< Ψ | ~r | Ψ > .

For a general function that can be expanded as F (r) = an rn , we define the average value as
P

Z
< F >= Ψ∗ (~r, t) F (r) Ψ(~r, t) =< Ψ | F | Ψ >

The spread, variance, or uncertainty in the observation of ~rx or F is defined in the root-mean
square way Z
(∆rx )2 = P (~r, t) [rx − < rx >]2 d~r =< rx2 > − < rx >2

and Z
2
(∆F ) = P (~r, t) [F − < F >]2 dr =< F 2 > − < F >2 .

3.4 Momentum space wave function


For a free particle wave packet, we have used the Fourier transform
~
ei(k·~r−ω(k)t) ~ ~
Z
Ψ(~r, t) = φ(k − k0 ) d3 r
(2π)3/2

to introduce the wave function in momentum space as Φ(~k, t) = e−iω(k)t φ(~k − ~k0 ). For that case
2
H0 Φ(~k, t) = ih̄ Φ̇(~k, t) yields h̄k ~
2m Φ(k, t) = ω(k) Φ.
More generally φ is not independent of time, but includes a change in the momentum components;
i.e., forces act. For the general case, we can write
~
eik·~r
Z
Ψ(~r, t) = Φ(~k, t) d~k
(2π)3/2

and identify Φ(~k, t) as the wave function in momentum space. Note that Φ is properly normalized
~ ~ 0
e−i(k−k )·~r 3 3 3 0 ∗ ~
Z Z
| Ψ |2 d3 r = d r d k d k Φ (k, t) Φ(~k 0 , t)
(2π)2
Z
= | Φ |2 d3 k = 1,
CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 53

which is Parseval’s Theorem (obtained via use of Dirac’s δ function). Therefore, we can identify

P (~k, t) =| Φ(~k, t) |2

as the probability for finding an electron with momentum ~k within interval d3 k at time t.
Using this result, we have

Z Z ~
e−ik·~r ~ k0 ei~k 0 ·~r

∗ ~
< ~r >= ∗
Ψ (~r, t) ~r Ψ(~r, t) d~r = Φ (k, t) [ ] Φ(~k 0 , t) d~k dk
~ 0 dr
~
(2π)3/2 i (2π)3/2
Z
1~
= Φ∗ (~k, t) [ ∇ ~ ~0 0 ~ ~ 00
k0 δ(k − k )] Φ(k , t) dk dk
i
Now use
~ k0 δ(k − k 0 ) = −∇
∇ ~ k δ(k − k 0 ),
which holds provided Φ’s vanish at the surface at ∞. Note: this step is equivalent to integration by
parts and dropping surface terms. Thus this operator property is defined with respect to boundary
conditions. We obtain the operator for ~r in momentum space from
Z
... < ~r >= Φ∗ (~k, t)i∇
~ k Φ(~k, t) d~k

Therefore, we see that in the ~r − representation Ψ(~r, t), ~r = a number, but in the p~ − representation, Φ(~k, t),
we must use ~r as an operator given by
~ k = ih̄∇
~r = i∇ ~p

Later, we will express this result as:

(~r 0 | ~rop | ~r) ≡ ~rδ(~r 0 − ~r)

and
(~k 0 | ~rop | ~k) ≡ i∇k δ(~k 0 − ~k)
~ p.
or ~r → ih̄∇
For a general function, F (r) = an rn , we get
P

Z
< F (r) >= Φ∗ (~k, t) F (i∇
~ k ) Φ(~k, t) d~k

or more generally we have


Z
< F (r) >= Ψ∗ (~r, t) F (r) Ψ(~r, t) d3 r

~ ~0
e−ik·~r eik ·~r
Z
∗ ~
= Φ (k, t) F (r) d3
r Φ(~k, t) d3 k d3 k 0
(2π)3/2 (2π)3/2
Z
= Φ∗ (~k, t) (~k | F | ~k 0 ) Φ(~k 0 , t) d~k d~k 0

with the definition


~ ~0
e−ik·~r eik ·~r
Z
(~k | F | ~k 0 ) = 3/2
F (r) d~r
(2π) (2π)3/2
for momentum space matrix elements.
If F(r) is to represent an observable, we must have it real < F >∗ =< F > which requires that
(~k | F | ~k 0 )∗ = (~k 0 | F | ~k) – a “matrix” in the continuous and infinite ~k, ~k 0 matrix variables or
F ∗T = F = F † is an Hermitian operator.
For the simple case of ~r we express the Hermiticity or self-adjoint property as:

(~k | F | ~k 0 ) = (~k | ~r | ~k 0 ) = i∇
~ k0 δ(~k − ~k 0 ) = −i∇k δ(~k − ~k 0 ) = (~k 0 | F | ~k)∗
CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 54

We are here practicing our use of the Dirac delta function in preparation for the introduction of
Dirac’s notation. For example, we will later use the steps:
Z Z
(k | ~r | Ψ) = (k | ~r | k )(k | Ψ) dk = i∇k δ(~k − ~k 0 )Φ(~k 0 )d~k 0 = i∇
~ ~ ~ 0 ~0 ~ 0 ~ ~ k Φ(~k) .

Having discussed the expectation value of ~rx , and general functions F (r), let us now turn to p~ and
general functions g(p). The corresponding steps are to introduce
Z
< p~ >= Φ∗ (~k, t) p~ Φ(~k, t) d~k

~ ~ 0
ei·k·~r e−ik·~r
Z
= 3/2
Ψ∗ (~r, t) d~r p~ Ψ(~r 0 , t) d~r 0 d~k
(2π) (2π)3/2
Z
~ r0 δ(~r − ~r 0 )] Ψ(~r 0 , t) d~r 0
= h̄ Ψ∗ (~r 0 , t) d~r [ i∇

Now, by integrating by parts ,or by using the delta function property


~ r0 δ(~r − ~r 0 ) = −i∇
i∇ ~ r δ(~r − ~r 0 ),

we have (dropping surface terms again)


Z
h̄ ~
< p~ >= Ψ∗ (~r 0 , t) ∇ Ψ(~r 0 , t) d~r =< Ψ | p~ | Ψ >
i
~ an operator in the r–representation.
where we associate p~ → h̄i ∇,
A general function g(p) has the corresponding property
Z
< g(p) >= Φ∗ (~k, t) g(p) Φ(~k, t) dk
~

Z
= Ψ∗ (~r, t) (~r | g | ~r 0 ) Ψ(~r 0 , t) d~r d~r0

~ ~0 0
eik·~r e−ik ·~r
Z
h~
0
(~r | g(p) | ~r ) = g(p) d~k = g( ∇ r − ~r 0 )
r )δ(~
(2π)3/2 (2π)3/2 i
or Z
h̄ ~
< g(p) >= Ψ∗ (~r, t) g( ∇ r ) Ψ(~r, t) d~r.
i
The requirement that the observable be real once again leads us to the idea of Hermitian “ma-
trices”
< g >=< g >∗
or
(~r | g | ~r 0 )∗ = (~r 0 | g | ~r ),
therefore, g ∗ = g T or g † = g which means explicitly that

(~r 0 | g | ~r)∗ = (~r | g | ~r 0 )

for any observable.


For the case of (~r | p~ | ~r 0 ) = h̄i ∇r δ(~r − ~r 0 ) this hermiticity property follows from

h̄ ~ h̄ ~
(~r | p~ | ~r 0 )∗ = − ∇ r − ~r 0 ) = ∇
r δ(~ r − ~r 0 ) = (~r 0 | p~ | ~r ).
r 0 δ(~
i i
Alternately, one sees that by integration by parts with zero surface terms
Z Z Z
∗ h̄ ~ h̄ ~ h̄
Ψ ∇ Ψ d~r = − (∇Ψ )Ψ d~r = ( ∇Ψ)∗ Ψ d~r.

i i i
CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 55

This last step of integrating by parts can be generalized to define the adjoint of any operator Ω. The
adjoint operator denoted by Ω† is defined by the following integration procedure
Z Z
Ψ1 Ω Ψ2 dτ = (Ω† Ψ1 )∗ Ψ2 dτ.

Here Ω† is the Adjoint operator. If Ω† = Ω, we have a self-adjoint or hermitian operator.


Written out in detail these steps are:
Z Z
Ψ∗1 (~r, t)(~r | Ω | ~r 0 ) Ψ2 (~r 0 , t) d~r d~r 0 = [ (~r 0 | Ω† | ~r )Ψ1 (~r, t) ]∗ Ψ2 (~r 0 , t)d~r d~r 0

Z
= Ψ∗1 (~r, t) (~r 0 | Ω† | ~r )∗ Ψ2 (~r0 , t) d~r d~r 0

which is equivalent to
(~r 0 | Ω† | ~r) = (~r | Ω | ~r 0 )∗
(the adjoint is seen here to be a transpose and a complex conjugation in the ~r, ~r 0 space ).
We can also represent our operators in the differential forms found earlier; namely, by

(~r | F (r) | ~r 0 ) = F (r)δ(~r − ~r 0 )


h~
(~r | g(p) | ~r 0 ) = g( ∇ r − ~r 0 ).
r )δ(~
i
Later functions of ~r and p~ will also be discussed.
The adjoint property can also be defined using the momentum space wave function. The steps
are Z Z
Φ∗1 Ω Φ2 d3 k = (Ω† Φ1 )∗ Φ2 d3 k
or Z Z
Φ∗1 (~k, t)(~k | Ω | ~k 0 )Φ2 (~k 0 , t)d~k d~k 0 = (~k 0 | Ω† | ~k)∗ Φ∗1 (~k, t) Φ2 (~k 0 , t) dk ~0
~ dk

... (~k 0 | Ω† | ~k) = (~k | Ω | ~k 0 )∗


which also defines Ω† .
Our case of
(~k 0 | ~r | ~k) = i∇
~ k0 δ(~k 0 − ~k) = (~k | ~r | ~k 0 )∗ = (~k 0 | r† | ~k)

shows that ~r is self-adjoint; Hermitian operators are defined to be self-adjoint; i.e. Ω† = Ω.


~ r δ(~r − ~r 0 ) = − h̄ ∇
Also our case of (~r | p~ | ~r 0 ) = h̄i ∇ ~ 0 r − ~r 0 ) = (~r 0 | p~ | ~r )∗ = (~r | p† | ~r 0 ),
i r δ(~
showed that ~r is self-adjoint or Hermitian. These are therefore respectable observables. Note the
integration by parts had already demonstrated that
Z Z Z

Ψ p~ Ψ dτ = (~ p Ψ) Ψ dτ ≡ Ψ∗ p~† Ψ dτ ... p~ = p~† .

CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 56

3.5 Expectation Values and Commutators


The expectation values of F (r) and g(p) type observables have been discussed and an improved
notation has been introduced. It is also necessary to consider functions of both ~r and p~, which we
denote as Ω(~r, p~). For example, the observables L ~ = ~r × p~, and p~ · A ~ are such observables. A velocity
dependent potential of the form (~ pV (r)~
p), is another example. For such functions the expectation
value is naturally given by
Z
h̄ ~
< Ω(~r, p~) > = Ψ∗ (~r, t) Ω(~r, ∇ r ) Ψ(~
r, t) d~r (3.1)
i
Z
= Φ∗ (~k, t) Ω(h̄i ∇~ p , p~) Φ(~k, t))d~k (3.2)

≡ < Ψ(~r, t) | Ω | Ψ(~r, t) > . (3.3)

~ r ) δ(~r − ~r 0 ) and (~k | Ω | ~k 0 ) = Ω(i ∇


Note: (~r | Ω | ~r 0 ) = Ω(~r, h̄i ∇ ~ k , h̄~k) δ(~k − ~k 0 ) are the differential
operator forms.
There is however an ambiguity in the definition of observables of such mixed form, which requires
us to be careful about the order of the ~r and p~ operators. For example, rx px is not equivalent to
px rx , in fact
h̄ ∂
ri pj Ψ = ri ( Ψ),
i ∂rj
whereas
h̄ ∂ h̄ h̄ ∂
pj ri Ψ = ( ri Ψ) = ( δi,j + ri )Ψ.
i ∂rj i i ∂rj
Thus we have the basic commutator property

pj ri − ri pj ≡ [pj , ri ] = δi,j .
i
Note that the above commutator also follows using the momentum space version
∂ ∂ h̄
pj ri − ri pj = pj ih̄ − ih̄ pj = −ih̄δi,j = δi,j .
∂pi ∂pi i
The commutator is therefore an operator equation, i.e. it is independent of representation.
Therefore, in defining observables of the form Ω(~r, p~) we must pay attention to the order of the
~r and p~ operators. As before, the restriction for observables is that the expectation value be real,
from which it follows that Ω(~r, p~) be self-adjoint. Let us determine the adjoint operator in some
simple cases and learn how to properly form Hermitian or self-adjoint operators.

3.6 Commutators
Our beginning point was to construct initial states in the form of wave packets which incorporate
the restrictions on our ability to measure position and momentum simultaneously. That idea led us
to Fourier transforms and the representation of general probability amplitudes as
Z
Ψ(~r, t) =< ~r | Ψ(t) >= < ~r | ~k >< ~k | Ψ(t) > d~k.

Our study of expectation values led us to this new notation and to identifying Hermitian operators
with the basic observables; for example, those given in the following table.√
Thus we find the commutator property [pk , rj ] = h̄i δk,j , where i ≡ −1, and the subscripts
k and j denote 1, 2, 3 or equivalently the x, y, z Cartesian components. This commutator, which
is independent of representation, is an outgrowth of our requirement to build a theory upon the
uncertainty principle. Later we shall indeed see that our inability to measure px and x simultaneously
with arbitrary precision, or that acts of measurement do not commute - the order of measurement
matters, will be a natural consequence of the operator algebra we now confront.
For now, let us develop the following points:
CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 57

Observable ~r − representation p~ − representation

~r ~r ~p
ih̄∇

h̄ ~
p~ i ∇r p~

Wavefunction < ~r | Ψ > < p~ | Ψ >

Table 3.1: Basic Observables in Coordinate and Momentum Space

(a) The operator equations are similar to those found classically in Poisson bracket form.
(b) The potential or forces acting on particles can then be defined for a quantum system in a
way that satisfies the correspondence principle; we obtain the classical limit and Ehrenfest’s
theorem.
(c) The uncertainty principle is a direct consequence of the noncommutivity of px and x or of any
such canonically conjugate variables.

In defining observables of the form Ω(~r, p~) we must pay attention to the order of the ~r and p~
operators. As before, the restriction for observables is that the expectation value be real from which
it follows that Ω(~r, p~) be self adjoint. Let us determine the adjoint operator in some simple cases
and learn how to properly form Hermitian or self-adjust operators Ω(~r, p~).
First consider Ω = xpx and determine Ω† . The steps are:

(x | xpx | x0 ) = x ∇x δ(x − x0 )
i
... (x | (xpx )† | x0 ) = (x0 | (xpx ) | x)∗

= −x0 ∇x0 δ(x0 − x)
i
0 h̄
= +x ∇x δ(x − x0 )
i
= (x | px x | x0 ) (3.4)
(3.5)

Therefore , we find (xpx )† = p†x x† = px x. Similar steps show that (px x)† = x† p†x = x px . Therefore,
neither xpx nor px x are self-adjoint; however, (xpx + px x)† = (xpx + px x)is self-adjoint and therefore
could represent an observable. Note that [(xpx − px x)/i]† = (xpx − px x)/i is also self-adjoint, but
not a classical “observable”, it goes to zero as h̄ → 0.
The above rule for taking adjoints can be generalized to

(Ω1 Ω2 )† = Ω†2 Ω†1 → Ω2 Ω1 ;

the last step holding for Hermitian operators Ω†1 = Ω1 , and Ω†2 = Ω2 . Prove of these statements is as
follows. The adjoint of a general product operator Ω ≡ Ω1 Ω2 is defined by transposing and complex
conjugating

(~r | Ω† | ~r 0 ) = (~r 0 | Ω | ~r )∗
Z
= (~r 0 | Ω1 Ω2 | ~r )∗ = d~r 00 (~r 0 | Ω1 | ~r 00 )∗ (~r 00
| Ω2 | ~r )∗
Z
= d~r 00 (~r 0 | Ω†2 | ~r 00 )(~r 00 | Ω†1 | ~r)

= (~r 0 | Ω†2 Ω†1 | ~r) (3.6)


CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 58

Hence (Ω1 Ω2 )† = Ω†2 Ω†1 . QED. These steps parallel those for ordinary finite dimensional discrete
matrices.
Therefore, we can building Hermitian operators using (Ω1 Ω2 + Ω2 Ω1 )/2 if Ω1 and Ω2 are Her-
mitian. On the other hand the commutator,(Ω1 Ω2 − Ω2 Ω1 ) is not Hermitian, but is anti-Hermitian.
Other properties of adjoint operators are: (Ω1 + Ω2 )† = Ω†1 + Ω†2 , and (a Ω1 + b Ω2 )† = a∗ Ω†1 +
b∗ Ω†2 .
These rules can be used to construct Hermitian functions of ~r and p~. However, the proce-
dure is not unique and some additional input is required. For example, from x2 and px we can
build (x2 px + px x2 )/2 as an Hermitian operator, but (x px x) is another different possibility, as is
a (x2 px + px x2 )/2 + b (x px x). We use the classical limit as a guide, but often the decision depends
on predictions using a hypothetical operator and experimental validation.

3.7 Dirac Notation


At this stage, we can make use of the Dirac delta function to introduce Dirac notation. The advantage
provided by the delta function is that operators and wave functions (or states) can be represented in
a “matrix” form in the continuous variables r and/or k. Thus one deals with a linear abstract vector
space to represent states, which can be thought of as the continuous and infinite limit of a finite
space matrix formulation - i.e. we shall introduce the idea of Hilbert space, the space of complex
bounded functions. Later, the fundamental notions of Hilbert space and operators will be discussed
more fully. For now let us simply begin to familiarize ourselves with some useful aspects of the Dirac
notation.
We have already seen how operators can be represented as “matrix-like” operations in the vari-
ables r and r0 by means of the delta function
Z Z
h̄ h̄
(r | Ω | r0 )Ψ(r0 )dr0 = ∇r δ(r − r0 )Ψ(r0 )dr0 = ∇r Ψ(r),
i i
where the integration over r0 replaces the sum in the matrix multiplication. Also, the idea of
Hermitian adjoint operators which is defined by the integral condition
Z Z
(Ω Ψ1 ) Ψ2 dr = Ψ∗1 ΩΨ2 dr,
† ∗

has been shown to be equivalent to a “matrix-like” definition

(r | Ω† | r0 ) = (r0 | Ω | r)∗

or simply Ω† = ΩT ∗ .
A self-adjoint or Hermitian operator is similarly given in“matrix” form as

(r | Ω† | r0 ) = (r0 | Ω | r)∗ ≡ (r | Ω | r0 ),

or Ω† = Ω.
Similar expressions can be given in momentum space representation.
Having introduced operators as continuous ∞ × ∞ matrices, we can now introduce Dirac’s “bra-
ket” notation which represents Ψ(r, t) and Ψ∗ (r, t ), as ∞ × 1 column and 1 × ∞ row vectors. We
have

Ψ(~r, t) ≡ (~r | Ψ(t) >


Ψ∗ (~r, t) ≡ < Ψ(t) | ~r ), (3.7)

where | Ψ(t) > denotes the state vector in an abstract vector space and we have taken its projection
on to the r- representation to find the wave function. The state < Ψ(t) | is similarly defined in
the adjoint vector space with < Ψ(t) |† ≡| Ψ(t) > being understood as the transformation of the
row state (the bra) to the column form-the ket. The vector or state can also be projected onto
momentum space and we define the momentum space wave functions as

Φ(~k, t) ≡ (~k | Ψ(t) >


Φ∗ (~k, t) ≡ < Ψ(t) | ~k), (3.8)
CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 59

Using this notation the Fourier transforms relating Φ(~k, t) and Ψ(~r, t) can be written concisely
as:
~
eik·~r
Z
Ψ(~r, t) = Φ(~k, t)d~k ⇒
(2π)3/2
Z
< ~r | Ψ(t) > = < ~r | ~k > d~k < ~k | Ψ(t) >;
~
e−ik·~r
Z
Φ(~k, t) = Ψ(~r, t)d~r ⇒
(2π)3/2
Z
< ~k | Ψ(t) > = < ~k | ~r > d~r < ~r | Ψ(t) > . (3.9)

The “transformation” factors for k − space ↔ r − space < ~k | ~r >, and < ~r | ~k >, have been defined
by
~
eik·~r
< ~r | ~k > =
(2π)3/2
~
e−ik·~r
< ~k | ~r > =
(2π)3/2
with < ~r | ~k > = < ~k | ~r >∗ . (3.10)

These expressions are really special cases of our previous case of < ~r | Ψ(t) >, where | Ψ(t) > is
taken as an eigenstate of momentum | Ψ(t) >→| ~k >, with p~op | ~k >= h̄~k | ~k > or

< ~r | p~op | ~k > = h̄~k < ~r | ~k >


Z
= d~r 0 < ~r | p~op | ~r 0 >< ~r 0 | ~k >

= ∇r < ~r | ~k >
i
~
eik·~r
... < ~r | ~k > = (3.11)
(2π)3/2

The factor (2π)3/2 provides the delta function normalization of these “plane waves”
Z
~ ~ 0
< k | k >= < ~k | ~r 0 > d~r 0 < ~r 0 | ~k 0 >= δ(~k − ~k 0 ),

and Z
< ~r | ~r 0 >= < ~r | ~k 0 > d~k 0 < ~k 0 | ~r 0 >= δ(~r − ~r 0 ).

These expressions are merely abbreviated means of expressing Fourier transformations and their
inverse. An even more formal way to express these relations is
Z
| ~r > d~r < ~r | = 1,

and Z
| ~k > d~k < ~k | = 1,

where 1 represents a unit operator in Hilbert space < ~r | 1 | ~r 0 >= δ(~r−~r 0 ), < ~k | 1 | ~k 0 >= δ(~k−~k 0 ).
For some more examples of Dirac Notation see Home work.
Later we shall return to discuss the Dirac notation and the idea of Hilbert space and shall then
see how Dirac’s formulation provides a neat way of expressing the equivalence of the wave and matrix
mechanics formulations of quantum mechanics. (The problem of quantum mechanics is thus reduced
to a problem in operator algebra).
CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 60

3.8 Operator Equations


From the basic commutator [pk , qj ] = h̄i δkj , we are able to deduce some general commutator prop-
erties and some operator equations that can be readily identified with classical equations of motion
in the correspondence principle limit.
First note that for noncommuting operators Ω1 , Ω2 and Ω3 , we have the obvious properties:

[Ω1 , Ω2 ] ≡ Ω1 Ω2 − Ω2 Ω1
[Ω1 , Ω2 + Ω3 ] = [Ω1 , Ω2 ] + [Ω1 , Ω3 ]
[Ω1 , Ω2 Ω3 ] = [Ω1 , Ω2 ] Ω3 + Ω2 [Ω1 , Ω3 ] . (3.12)

Furthermore, from the basic commutator [pk , qj ] = h̄i δkj we obtain the following generalizations:

∂ n
[rx , pnx ] = ih̄npxn−1 = ih̄( p )
∂px x
[rx , pny ] = [rx , pnz ] = 0 etc.
... [~r, Ω(~
p )] = ih̄(∇~ p Ω(~ p ) ). (3.13)

~ p is an operator
This generalization follows immediately using the p-representation for which ~r → ih̄∇
and p~ is simply a numerical vector. Similarly, we have:
∂ n
[px , rxn ] = ih̄nrxn−1 = ih̄( r )
∂rx x
[px , ryn ] = [px , rzn ] = 0 etc.
... [~
p, Ω(~r )] ~ r Ω(~r ) ).
= ih̄(∇ (3.14)

We can now consider functions of r and p simultaneously - called Ω(~r, p~). When referring to a
Hamiltonian, we shall call the general function Ω = H(~r, p~). It follows that

h̄ ~
[~
p, Ω(~r, p~)] = (∇r Ω(~r, p~) )
i
[~r, Ω(~r, p~)] ~ p Ω(~r, p~) ).
= ih̄(∇ (3.15)

Although these are seen from using the r and then the p representations, these results are represen-
tation independent. They can be derived using Equation 3.12
We can apply these general commutators or operator equations to our free particle Hamiltonian

~p p2 p~
[~r, H0 ] = ih̄∇ = ih̄( )
2m m
or
p~ 1 “d~r ”
= [~r, H0 ] → = ~vparticle .
m ih̄ dt
The last part of the above statement refers to later steps which will enable us to define an time
derivative operator “d~
r”
dt , which is used to construct the particle velocity. Here we see a first approach
to quantum equations that can be of “classical form.” The case of
2
~ r p = 0,
[~r, H0 ] = ih̄∇
2m
is a form easily understood to be related to the classical limit of no momentum or kinetic energy
change for a free particle.

3.9 Time Derivative Operator


The time development of the quantum system has been described by having the wave function be
a function of t determined by HΨ = ih̄Ψ̇. We therefore need to consider the expectation value of a
CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 61

general operator Ω before we can discuss the time evolution of our observables. Note the description
is being presented now in the Schrödinger-picture, i.e. the time dependence is in the wave function.
Alternate descriptions will be discussed later (i.e. the Heisenberg and Interaction or Dirac pictures).
The expectation value is given in general by
Z
< Ω >= Ψ∗ (~r, t)(~r | Ω(t) | ~r 0 )Ψ(~r 0 , t)d~r d~r 0 .

In addition to the time dependence given by the wave function, we allow for the possibility of some
explicit dependence of Ω(t) on time ((~r | Ω(t) | ~r 0 ) corresponding to some external driving effect.
Thus, the time derivative is
Z
d<Ω> ∂
= ( Ψ∗ (~r, t))(~r | Ω(t) | ~r 0 )Ψ(~r 0 , t)d~r d~r 0
dt ∂t
Z

+ Ψ∗ (~r, t)(~r | Ω(t) | ~r 0 )( Ψ(~r 0 , t))d~r d~r 0
∂t
Z

+ Ψ∗ (~r, t)(~r | ( Ω(t)) | ~r 0 )Ψ(~r 0 , t)d~r d~r 0 . (3.16)
∂t
Since Z
ih̄Ψ̇ = HΨ = (~r | H(t) | ~r 0 )Ψ(~r 0 )d~r 0 ,
2
with for example (~r | H0 | ~r 0 ) = − 2m

∇2r δ(~r − ~r 0 ). and for a local, central potential (~r | V | ~r 0 ) =
V (r)δ(~r − ~r ), and H is Hermitian (~r | H | ~r 0 )∗ = (~r 0 | H | ~r ), we find
0

d<Ω> ∂ 1
=< Ω(t) + [Ω, H] > (3.17)
dt ∂t ih̄
This suggests b that the “time derivative operator” is given by
“dΩ” ∂ 1
= Ω(t) + [Ω, H]. (3.18)
dt ∂t ih̄
(Here we see the Hamiltonian’s role as a generator of time development.) The above operator is to
be understood, as derived, to be the operator whose expectation value yields the time derivative of
the expectation value < Ω >, i.e.
d<Ω> “dΩ”
=< >!
dt dt
We can apply this general result to some examples.
The operator ~r, which is simply a number in the r-representation, has no explicit dependence on
time; nevertheless it has a time derivative operator given by

“d~r ” 1
= [~r, H].
dt ih̄
p2
For a free particle case H → H0 = 2m , the above result gives

“d~r ” 1 p~
= [~r, H0 ] = = ~vparticle .
dt ih̄ m
We have succeeded in obtaining an operator which can be used to get

d < ~r > “dΩ ” < p~ >


=< >= = ~vparticle .
dt dt m
The classical result < ~r >= ~vparticle t. has just been deduced.
b To generalize from an expectation value result to an operator equation, it is necessary to show that all “matrix

elements” satisfy the same equation. That is, the operator equation is more general than the result for just one set of
matrix element; which is the basic idea of an operator in abstract space. In this case, the same steps show that the
time development result is the same for different states < Ψ | Ω(t) | Ψ0 >,i.e., for all matrix elements.
CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 62

Using p~ in our general time derivative result, the corresponding time derivative operator is
“d~
p” 1
= [~ p, H].
dt ih̄
Thus we verify that for free particles
“d~
p” 1
= [~ p, H0 ] = 0,
dt ih̄
d<~
p>
dt =< “d~ p”
dt >= 0; we get zero Force.
There are special cases of Ehrenfest’s theorem. Here, we use the classical limit obtained above
as a guide to how one should introduce potentials or forces - Thus the Ehrenfest theorem is here
used as a restriction; namely, that the classical limit yield Newton’s laws of motion for the average
p2
values. The introduction of potentials using H = 2m + V (r), clearly satisfies this correspondence
principle condition since we have
d~r 1 p~
= [~r, H] = ,
dt ih̄ m
and
d~
p 1 ~ r V (r).
= [~r, H] = −∇
dt ih̄
Using the definition of the time derivative operator discussed earlier, we see that
d < ~r > < p~ >
= ,
dt m
or < p~ >= m~v , and
d < p~ > ~ r V (r) >= F~ .
== − < ∇
dt
or F~ =< p~˙ > .
Newton’s law thus follows from the definition of average values and the introduction of potential
as H = H0 + V. This is Ehrenfest’s theorem, which is actually used here as a condition that tells us
to introduce V so as to assure obtaining the classical limit.
More needs to be said to establish the above classical limit. For example, one needs to show that
~ r V (r) >=
the force is really a classical function in the sense F (< r >, t). We need to show that < ∇
~
∇r < V (r) > in the classical limit. This is discussed in Gottfried.
Instead of following his discussion, I shall complement his treatment by obtaining the classical
limit by considering the relationship between commutators and Poisson Brackets.

Conservation of Energy
We can pick Ω = H and apply our result for the time derivative operator
d<H> ∂H
=< >,
dt ∂t
which is d<H>
dt = 0, in the absence of an explicit dependence on time. An explicit dependence on
time would mean that energy is pumped into or out of the system. Hence, with no such outside
effect, we get our conservation of energy law < H >=average energy=constant. Much more will be
said about such conservation laws and related symmetries later.

3.10 Commutators and Poisson Brackets


c
One way of obtaining the classical limit from quantum mechanics is to consider the h̄ → 0 limit of
the following equations:
1
[rk , pj ] = δkj
ih̄
c This is a trick for getting the correspondence principle or classical limit, by making the quantum action negligible.

It is often simpler than taking the action to be large compared to h̄.


CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 63

dΩ ∂ 1
= Ω(t) + [Ω, H]
dt ∂t ih̄
1 ~
[~r, H] = (∇p H),
ih̄
1 ~ r H).
p, H] = −(∇
[~ (3.19)
ih̄
To take this limit let us consider a general function of the operators ~r and p~, which are expanded
about their average or classical values < ~r > and < p~ > in a Taylor series:
3f
X ∂Ω ∂Ω
Ω(~r, p~) = Ω(< ~r >, < p~ >) + [( ) Qi + ( ) Pi ] + O(h̄) + · · · , (3.20)
i=1
∂ < ri > ∂ < pi >

where
Qi = ri − < ri >
and
Pi = pi − < pi > .
The above case has been generalized for f particles, with 3f degrees of freedom.
Here higher order terms depend on higher powers of h̄ and will be neglected as h̄ → 0. In
evaluating the derivatives we also used the expansions
∂Ω ∂Ω
= + O(h̄) · · ·
∂pi ∂ p̄i
and
∂Ω ∂Ω
= + O(h̄) · · · ,
∂ri ∂ r̄i >
where the alternate notation for the averages < ri >= r̄i and < pi >= p̄i have been used. These
derivatives are classical quantities - classical derivatives! Since < ~ri > and < p~i > are just numbers,
we have the standard commutators holding forPi and Pi ,
[Qk , Pj ] = ih̄δkj [Qk , Qj ] = 0 [Pk , Pj ] = 0,
i.e. a simple canonical transformation is used.
Applying this expansion to operators Ω1 and Ω2 we obtain the expansion of the commutator:
3f
X ∂Ω1 ∂Ω2 ∂Ω1 ∂Ω2
[Ω1 , Ω2 ] = [( )( )−( )( )] [Qi , Pj ] + O(h̄) + · · · (3.21)
i,j=1
∂ < ri > ∂ < pj > ∂ < pj > ∂ < ri >
or
1
[Ω1 , Ω2 ] = {Ω1 , Ω2 }P B .
lim (3.22)
ih̄h̄→0
Here the Poisson Bracket (PB) is defined by
3f
X ∂A ∂B ∂A ∂B
{A, B}P B ≡ − , (3.23)
i=1
∂ri ∂pi ∂pi ∂ri

which involves classical derivatives only.


Now apply this result to our various commutators and find:
[rk , pj ] X ∂ r̄k ∂ p̄j
lim = {r̄k , p̄j }P B = j0 = δkj
h̄→0 ıh̄ ∂ r̄j 0 ∂ p̄j 0
dΩ ∂Ω(r̄, p̄)
lim = + {Ω(r̄, p̄), H(r̄, p̄)}P B
h̄→0 dt ∂t
1
lim [r, H] = {r̄, H(r̄, p̄)}P B = ∇p̄ H(r̄, p̄)
h̄→0 ih̄
1
lim [p, H] = {p̄, H(r̄, p̄)}P B = −∇r̄ H(r̄, p̄), (3.24)
h̄→0 ih̄

where it is understood that classical quantities (r̄ ≡< r >, p̄ ≡< p >) (expectation values) are
used within the Poisson Bracket terms. We shall now identify these classical limits as the classical
equations of motion in the Hamiltonian formulation.
CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 64

3.11 Review of Classical Mechanics


To make contact with the classical limit of quantum mechanics, it is useful to quickly review the
basic formulations of classical mechanics. Of course, we have Newton’s law

F~ = p~˙ = m~¨r

and the Lagrange equations of motion based on the action principle, which expresses dynamics in
terms of generalized coordinates (r, θ, φ for example). The Lagrangian

L = L(q, q̇, t) ⇒ T (q̇) − V (q)

is used in the action principle Z t2


δ L(q, q̇, t)dt = 0
t1

with δL(t1 ) = δL(t2 ) = 0 to derive


d ∂L ∂L
− =0
dt ∂ q̇ ∂q
This formulations provides the great practical convenience of setting up the differential equations
d
(2nd order, as in dt p~ = −∇V~ ) using coordinates that are selected to simplify the problem by
accounting for constraints. The canonically conjugate momenta are defined by
∂L
pi =
∂ q˙i
also called generalized momenta. Often L displays a symmetry property such as being independent
of a generalized coordinate q ; it then follows that p is independent of time - it is conserved. Thus
symmetries → conservation laws, which simplify our treatment of dynamics. (For example, if L is
independent of θ and φ, then orbital angular momentum is conserved.) The main advantage of the
Lagrangian formulation is to use such symmetries and generalized coordinates to obtain equations
of motion for rather complicated systems, such as coupled oscillators.
The Hamiltonian formulation is based on using the variables q and p - canonically conjugate
variables - to obtain a pair of first order coupled equations involving a function H

H = H(p, q, t) = q̇ p − L(q, q̇, t).


2
p
The Hamiltonian, H, is often of the form 2m + V (q). Note that the action is related to L and to p
and q in a simple way for the case of H = constant = energy; namely
Z Z
Ldt = −H∆t + p dq.
R
Therefore, the action principle becomes δ p dq = 0. Hamilton’s equations are obtained from the
above definition to be
∂H
= q̇
∂p
and
∂H ∂V
= = −F = −ṗ
∂q ∂q
∂H
= −ṗ
∂q
which are two equations first order in time derivations. The original motivation for obtaining such
equations was to explore the analogy between ray optics and particle motion - Fermat’s principle
and least action. The Hamiltonian provided a formal apparatus - in the Hamilton-Jacobi equations
- for understanding that analogy. It did not provide much practical help in problem solving, until
the advent of quantum theory. Another aspect of Hamilton’s equations is the use of canonical trans-
formations and their generators to describe the relationship between symmetries and conservation
principles. That aspect comes to full fruition in the quantum theory. I will discuss these topics later.
CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 65

Our present interest is to describe the Hamilton equations in terms of Poisson brackets. The
time dependence of any classical function of p and q, Ω(p, q) is expressed as
f
d ∂Ω X ∂Ω ∂qi ∂Ω ∂pi
Ω(q, p, t) = + +
dt ∂t i=1
∂qi ∂t ∂pi ∂t

f
∂Ω X ∂Ω ∂H ∂Ω ∂H
= + −
∂t i=1
∂qi ∂pi ∂pi ∂qi

∂Ω
= + {Ω, H}P B
∂t
where we have introduced the Poisson bracket defined by
f
X ∂A ∂B ∂B ∂A
{A, B}P B = ( − )
i=1
∂qi ∂pi ∂qi ∂pi

for f degrees of freedom. The following properties follow immediately from this definition

{pi , pj }P B = {qi , qj }P B = 0

{pi , qj }P B = −δij
{B, A}P B = −{A, B}P B
Note that the Poisson bracket provide a “canonically invariant” means of expressing the Hamilton
equations of motion since the above time derivatives reduce to
dq ∂H dpi ∂H
= {qi , H} = and = {pi , H} = −
dt ∂pi dt ∂qi
i.e. Hamilton’s equation. We omit the proof of the invariance of these equations under a canonical
transformation (see Goldstein’s Mechanics). We thus see that the h̄ → 0 limit of our commutators
should lead to the classical equations of motion written using Poisson brackets.
CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 66

3.12 The Uncertainty Principle


In the limit of large actions (or by taking h̄ → 0) our quantum mechanical operator equations reduce
to the Hamilton equations of classical physics expressed in Poisson Bracket form. The symmetry
principles, canonical transformation and general framework of quantum mechanics will be seen later
to be closely related to that of classical physics. It is simply the introduction of h̄, the related
introduction of operators, and the probabilistic interpretation that are basic to quantum theory
alone.
We now shall see that the introduction of h̄ and of operators using the basic commutator
[ri , pj ] = h̄i δij
is simply a statement of the uncertainty principle. To see that we must, of course, also define the
idea of the expectation value and the probability. The fact that the operator equations stores the
uncertainty principle requirements in our dynamics should be no surprise. We deduced the operator
equation starting from wave function and the Fourier transforms inserted into the definition of
expectation values. It was originally the Fourier transform that we used as a means of storing
the ∆p ∆x ≥ h̄ condition or the quantum description. To see the relationship between the basic
commutator and the uncertainty principle, let us first define shifted hermitian operators
Qi ≡ ri − < ri >
Pi ≡ pi − < pi >
< Qi >= 0 < Pi >= 0.
Where the <> brackets indicate average values. We have
[Qi , Pj ] = h̄i δij
and also (∆ri )2 =< Q2i >; (∆pi )2 =< Pi2 > . Consider
(Pi − iαQj ) Ψ >≡| α >,
where α is a real parameter and the vector |α > satisfies
<α|α>≥0
or in detail Z
[(Pi − iαQj )Ψ]∗ (Pi − iαQj )Ψ dτ ≥ 0
Z
Ψ∗ (Pi2 + α2 Q2j + iα[Qj , Pi ])Ψ dτ ≥ 0
Z
Ψ∗ [Pi† + iα∗ Q†j ][Pi − iαQj ]Ψdτ ≥ 0
or
< Pi2 > +α2 < Q2j > +iα < [Qj , Pi ] > ≥ 0
Recall P and Q are hermitian and α is a real parameter of any value. We can now write the above
as
(∆pi )2 + α2 (∆rj )2 − h̄α δij ≥ 0
h̄δij ∆pi 2
(∆rj )2 {α2 − 2
α+( ) }≥0
(∆rj ) ∆rj
h̄ 2 h̄2 δij ∆pi 2
(α − δ ij ) − +( ) ≥0
2(∆rj )2 4(∆rj )4 ∆rj

We now take α = δ
2∆rj2 ij
in which case to assure inequality or at least positive definite character of
right hand side we need
∆pi h̄
≥ δij
∆rj 2(∆rj )2
or

∆pi ∆rj ≥ δij QED
2
CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 67

3.12.1 Generalization to [A, B] = i C


It is of considerable interest to see how the general commutator

[A, B] = i C with A† = A, B † = B, C † = C,

also generates an uncertainty principle. The Hermitian operators can represent observables. Using
our previous steps, we have
A− < A >= A = A† ;
B− < B >= B = B †
[A, B] = [A, B] = iC
(∆A)2 ≡< A2 >=< A2 > − < A >2 .
(∆B)2 =< B 2 >=< B 2 > − < B >2 .
Consider: (A − iαB)Ψ and Z
Ψ∗ (A + iαB)(A − iαB)Ψdτ ≥ 0

< A2 > +α2 < B 2 > +iα < [B, A] >≥ 0


(∆A)2 + α2 (∆B)2 + α < C > ≥ 0
<C> ∆A 2
(∆B)2 {α2 + 2
α+( ) }≥0
(∆B) ∆B
<C> 2 < C >2 ∆A 2
(α + ) − +( ) ≥0
2(∆B)2 22 (∆B)4 ∆B
for any real α, say α = − < C > /2(∆B)2 ... we must have

∆A 2 < C >2
( ) ≥
∆B 4(∆B)4

and our final result


|< C >|
∆A ∆B ≥ .
2
or
|< i[B, A] >|
∆A ∆B ≥
2
Note that if C = 0, we obtain ∆A ∆B = 0; which is the important case of commuting observables.
In the case of such commuting operators, both of the associated observables can simultaneously
detected with precision !

Applications:
Consider, the equation of motion
d~
p
[~
p, H] = ih̄
dt
d
and identify A = p~x , B = H, and C = h̄ dt px . Applying the general result, we have immediately

h̄ dpx
∆px ∆E ≥ < >
2 dt
where ∆E =< H 2 > − < H > . A time interval appears as the ratio
∆px ∆px
d
= d
= time interval ≡ ∆t.
< dt px > dt < px >

Here ∆t is the uncertainty in the time that the particle experienced a change in momentum. We
have the important result that
∆E ∆t ≥ h̄/2.
CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 68

Also, from the relation, [~r, H] = ih̄ d~


r
dt , we see that

∆E ∆t ≥ h̄/2

with a time interval now defined by


∆rx
∆t ≡ d
,
dt < rx >
which is the uncertainty in time wave packet passed = ∆rx /vparticle .
Or for a general operator Ω, from [Ω, H] = ih̄Ω̇ we get

∆E ∆t ≥ h̄/2
d
with ∆t = ∆Ω/| dt < Ω > |!
d
It is important to do this right so we can interpret ∆t
< rx >
∆t = d<rx >
dt

as for example the uncertainty in time the wave packet passed by.
Note we shall later deal with other operator equations of the form [A, B] = i C –each of which
will yield an uncertainty principle. For example, we will find such relations for the electromagnetic
fields [Ax (~r, t), Ax (~r, t)] = · · · , from which we will get

∆Ex ∆By ≥ h̄/( ).

(Also, we will get more general relations such as ∆N ∆Φ ≥ 1 referring to the number of photons and
the phase of an electromagnetic “wave”).
Finally, the relation
[Lx , Ly ] = ih̄Lz
leads to
∆`x ∆`y
≥ h̄
< Lz >
and
[Lz , Lx ] = ih̄Ly
leads to
∆`x
∆`z ≥ h̄
< Ly >
or defining the ratios as angle uncertainties, we get for example

∆`z ∆φ ≥ h̄.

dA wrong - often used derivation is to note that



[ih̄ , t] = ih̄
∂t
and to conclude that
∆E∆t ≥ h̄
follows. This is wrong because “t” is not an operator but a parameter. We need to prove an uncertainty princi-
ple for canonically conjugate operators–independent of representation. We have no “time operator in the energy
representation.”
CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 69

3.12.2 Minimum Uncertainty Wave Function


Clearly the equality sign will apply if

(Pj − iαQj )Ψ = 0

and hence
∆pj ∆rj = h̄/2
Using this condition, we find

( ∇j − iαrj )Ψ = [< pj > −iα < rj >]Ψ,
i
whose solution is 2
i α
Ψ = N e h̄ <~p>·~r e− 2 (~r−<~r>)
Note:
h̄ h̄
pj Ψ = [< pj > + α(−1)(~rj − < ~rj >)]Ψ = ∇j Ψ
i i
or
[pj − ih̄ α rj ]Ψ = [< pj > −h̄iα < rj >]Ψ.
... Gaussian wave packet satisfies minimum uncertainty principle ∆p ∆x = h̄/2. Earlier, we indeed
found
h̄ ā h̄
∆rx (t) ∆px = = [1 + 42 ]1/2 ≥ h̄/2
2a 2
and at t = 0
∆rx (0)∆px = h̄/2 !
CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 70

3.13 The Classical Limit


(WKB approach)
Ref: Messiah Chapter VI
Schiff - p 269
Dirac
Goldstein Chapter 9

The classical limit has been obtained in two ways. The first method involved taking expectation
values of operator equations to obtain Ehrenfest’s theorem; i.e.,
d 1 ~
“ p~ ” = [~ p, H] = −∇V
dt ih̄
which yields
d
< p~ >= F~ = − < ∇V
~ >≈ −∇ ~ <V >.
dt
The second method consisted of expanding our operator equations about the classical or average
values. That approach yielded the theorem
1
lim [Ω1 , Ω2 ] = {Ω1 , Ω2 }P B .
h̄→0 ih̄
The operator equations were then seen to yield Hamilton’s equations of motion in Poisson bracket
form.
Now I shall discuss a third way of going to the classical limit. This method is less formal since
it makes use of our ability to visualize waves; the method is called the WKB (Wentzel-Kramers-
Brillouin) method and is useful for getting to a classical limit and also as a means of introducing
quantum effects approximately (especially for 1D problems, see later).
The WKB method is based on identifying the amplitude and phase functions in the wave functions
i
Ψ(~r, t) = A(~r, t) e h̄ S(~r, t) .

There is no loss of generality in taking both A and S to be real. [Note that since | ψ |2 ≡ 1/ volume |
A |2 has the units of 1/volume, and S(~r, t) has the units of h̄. That is, S has the units of action
(h̄), i.e. of E t and/or of p~ · ~r. In fact, for our wave packet we had S = h̄(~k0 · ~r − ω0 t) = p~0 · ~r − E0 t.
We can think of A (~r, t) as an envelope (packet) for an oscillating wave function, whose rate of
oscillation is determined by S(~r, t).
CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 71

Our earlier considerations have led us to conclude that the classical limit, for which negligible
spreading of a wave packet ball occurs, is obtained when
–B L
λ
≡ << 1.
a a
Here, a is the size (∆x) of A(~r, t), λ –B = k0−1 is the average deBroglie wave length and L is
the macroscopic distance travelled. (This expression for  follows from our wave packet analysis
or directly from the uncertainty principle). Since L is much greater than a, small spreading (or
classical motion) occurs when λ–B << a. We shall obtain this same result in our present discussion.
The packet for classical motion looks like:

<----------a------------>

Figure 3.1: A wavepacket with envelope A

Lots of oscillations corresponds to high kinetic energy, and high quantum numbers n; for example,
in harmonic oscillator we find

Ψ V= H12L m Ω2 r2 ®

V= H12L m Ω2 r2 ®

(a) (b)

Figure 3.2: Harmonic Oscillator with Low (a) and High (b) Quantum Number Wavefunctions
i
Let us now insert ψ = A e h̄ S into the continuity equation:
ρ =| A(r, t) |2 = A2 (r, t) = probability/ vol

~j = h̄ (ψ ∗ ∇ψ
~ − ψ ∇ψ~ ∗ ) = h̄ A2 (r, t) i ∇S
~ ∗2
2im 2im h̄
~
~j = ρ ∇S = ρ ~vef f
m
(recall A and S are real). Here, we get an effective velocity from the gradient of the action S
~
∇S
vef f ≡
m
This has the meaning shown in the figure which depicts that a “ray” or “particle” moves perpen-
dicular to the wave front, where the wave front is a surface of constant phase (of constant S), i.e.
CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 72

S(~r, t) = constant gives a moving wave front. For example, a plane wave has S = p~ · ~r − Et, ~vef f =
p
~
m = vparticle = ~vef f for S = p · r − Et = constant. Here we see the analogy QM/CM as wave /ray
optics !

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


.... .... ....
.... .... ....
... ... ...
... ... ...
... ... ...
... ... ...
.x- . .x-
... .. ..
... .... ....
.. .. .. vef f ~
= ∇S
.. .. ..
..... ..... .....
... ... ...
.... .... ....
Wavefronts S

Figure 3.3: Wave Fronts and a Ray

Now we can use this Ψ in the Schrödinger equation and obtain (see following section)
∂ ~ 2
(∇S) h̄2 ∇2 A
− S(~r, t) = +V −
∂t 2m 2m A
We also get
1 ~ ~ A 2
(∇A)(∇S) − ∇ S Ȧ = −
m 2m
~ ·~j = 0). As is expected (see i ∂ → continuity
but that is equivalent to the continuity equation (ρ̇ + ∇ ∂t
discussion.)
2 2
h̄ ∇ A
The classical limit can now be obtained by neglecting the term 2m A . Then we get the Hamilton
- Jacobi equation for the classical action S
∂S ~ r)
− (r, t) = H(∇S,
∂t
p2
where H = 2m + V (r).

Dirac’s Derivation
Dirac gives a very simple derivation as follows
i
ih̄ψ̇ = ih̄[Ȧ + ṠA]e+iS/h̄ = H(p, q) e+iS/h̄ A

or

− Ṡ] = e−iS/h̄ H(p, q) e+iS/h̄ = H(~
A[ih̄ p + (∇S), q)A,
A
where the last step follows from the properties:
e−iS/h̄ qe+iS/h̄ = q;
and
h̄ ~
e−iS/h̄ p e+iS/h̄ = e−iS/h̄ ∇e+iS/h̄ = p~ + ∇S,
i
~ These properties are applied to all powers of
as in a classical canonical transformation p̃ = p + ∇S.
q and p and hence to the function H(p, q). Now in the limit h̄ → 0 we get
h̄ Ȧ
∇ → 0 ih̄ → 0,
i A
and the Hamilton - Jacobi equation is obtained again
∂S ~
−Ṡ = − = H(∇S), q).
∂t
CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 73

Physical Meaning
The physical meaning of this classical result can be seen as follows:
∂S
− = E = H(mvef f , q)
∂t
or better taking the gradient
∂ ~ ∂
+ ∇S = −∇H = −∇V = Force = + mvef f = p~˙ .
∂t ∂t
or F = ma.

The case of − ∂t S = E corresponds to a “stationary state” or a state of fixed definite energy.
Et
Then [ψ ≡ e−i h̄ ψn (r)] and
S = −Et + W (r)
and the Hamilton - Jacobi equation becomes
~
E = H(∇W, q)

compare to Goldstein page 280.


We can introduce S = W (r) − Et into the original equation and find

∂S ~ )2
(∇W h̄2 ∇2 A
− =E= +V − ;
∂t 2m 2m A
neglecting the last term, we get
~ |=| ∇W
~
p
| ∇S |= 2m(E − V )

or
h̄ p
p(r) = h̄k(r) = – = 2m(E − V (r))
λ(r)
~ –(r) = h̄/ 2m(E − V (r)) is a local wave length, which varies with the
p
with p~(r) ≡ ∇S(r). Here λ
spatial dependence of the potential. When | E − V (r) | is large λ – → small and we have a chance
to get “classical motion”; however, it is clear that near the classical turning points (T.P.), the wave
length λ(r) → ∞ since E = V (rT.P. ) defines the turning point. For E >> V (r), which is the case
of having large quantum numbers, the wave length λ(r) → 0 is very small. Therefore, the classical
limit is obtained for high quantum numbers, as was the case in earlier versions of the correspondence
principle. Near the classical turning points, one does not get classical motion but the most quantum
mechanical situation. It is here (at the T.P.) that the quantum mechanical penetration effect occurs!
To gain further insight, and to relate the classical limit to the spreading idea, let us estimate the
contribution of the ∇2 A/A term. Now we have

~ 2 = 2m(E − V ) + h̄2 ∇2 A h̄2 2


–2 (r) ∇ A ].
(∇S) = 2 [1 + λ
A – (r)
λ A

The classical limit holds if


2
–2 (r) ∇ A << 1
λ
A
or (for A ∼
2
/a2
= e−r ) we get:
–2
λ
<< 1.
a2
a ∼ wave packet size.
... We recover our original condition for little spreading, mentioned at the beginning of this lecture;
namely, that λ–ef f << a. Hence, the classical limit is understood as arising for large quantum numbers,
large action, and equivalently for small spreading.
As a further connection with classical ideas consider a particle in a potential V (r), then we can
use
p dS
∇S = 2m(E − V ) = or
dr
CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 74
p
dS = 2m(E − V ) dr = dW
to write the action as Z q p
S = action ≡ −Et + 2m(E − V ) dr or
q0
Z q
= −Et + p(r) dr.
q0
Also note that, in a simplified version, for H = T + V
Z Z Z Z
Ldt = (T − V )dt = 2T dt − Hdt
Z
dt
= 2T dq − Et
dq
Z
= pdq − Et = action

... large S → large action. e


The final point is simply to write the action of the system as above and to justify our associating
S with the idea of action.

Some WKB Relations


i
Ψ = Ae h̄ S
h̄2 2
ih̄Ψ̇ = (− ∇ + V )Ψ
2m
i i i
ih̄{Ȧ e h̄ S + A Ṡ e h̄ S } = ih̄ψ̇

i
= {ih̄Ȧ − AṠ} e h̄ S
i
~ ∇A)e~ i i ~ i
∇2 ψ = ∇2 A e h̄ S = ∇{( h̄ S + A (∇S) e h̄ S }

2 i
S 2i ~ ~ i
S i i 1 i
= (∇ A)e h̄ + (∇A) · (∇S)e h̄ + A ∇ Se h̄ S − 2 A(∇S)2 e h̄ S
2
h̄ h̄ h̄
i 2i i 1
= e h̄ S {∇2 A + (∇A · ∇S) + A(∇2 S) − 2 A(∇S)2 }
h̄ h̄ h̄
h̄2 2 h̄i ~ ~ h̄i 1
[ih̄Ȧ − AṠ] = − ∇ A − (∇A)(∇S) − A(∇2 S) + A(∇S)
2m m 2m 2m
1 ~ ~ A 2
Ȧ = − (∇A)( ∇S) − ∇ S
m 2m
h̄2 2 1
−AṠ = − ∇ A+ A(∇S)2 + V A
2m 2m
~ 2
(∇S) h̄2 ∇2 A
−Ṡ = +V −
2m 2m A
d 2
Multiply above by 2mA and use 2mAȦ = m dt A to get
d 2 ~ 2 ∇S]
m A = −2A(∇A)(∇S) − A2 ∇2 S = −∇[A
dt
~
∇S
ρ = A2 ~j = A2 = ρ~vef f
m
... as before
~ · ~j = −ρ̇

No surprise since Schrödinger equation already includes continuity equation
d h̄2 ∇2 A
−Ṡ = − S = H(∇S, q) − 7−→ H(∇S, q)
dt 2m A h̄→0
where H(pq) = p2 /2m + V (q).
e 2T = 2 m 2
q = q̇!
q̇ q̇ 2
CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 75

3.14 Stationary States


In the last lecture, I discussed how the commutators are a formal means of storing the uncertainty
principle. We had generalized that remark to show that
1
[A, B] = iC → ∆A∆B ≥ |< C >|
2
where A and B are Hermitian operators (and hence, C + = C also).
A special case of interest is when two Hermitian operators commute [A, B] = 0 and hence
∆A ∆B = 0
Thus, commuting Hermitian operators represent observables which can be simultaneously mea-
sured to arbitrarily high precision. For example, we shall later find that the energy and angular
momentum (as well as the parity, `z · · ·) can be simultaneously measured and used to characterize
or label a state using quantum numbers. For now, let me just note that we would like to find a
complete set of commuting observables, also called compatible observables, to completely describe a
quantum state.

3.14.1 Eigenfunctions
As a step toward finding a complete description of a quantum state, let us introduce the idea of
stationary states defined by
i
Ψ(~r, t) = e− h̄ E t ψE ( ~r ) = e−iωt ψE ( ~r ).
For such states, the Schrödinger equation reduces to a time-independent form, provided H is not
dependent on time:

H Ψ(~r, t) = ih̄ Ψ(~r, t) → H ψ(~r ) = E ψ(~r).
∂t
This is an eigenvalue equation which has some important properties. These states are called sta-
tionary because the current and density have the following properties for real E:
ρ(~r, t) = Ψ∗ (~r, t)Ψ(~r, t) =| ψE (~r) |2 = ρE (~r)

~j(~r, t) = h̄ (ψ ∗ ∇ψ
~ − ψ ∇ψ
~ ∗ ) = ~jE (~r)
2im
i.e., density and current are time independent and ... ∇
~ · ~jE = 0 !
Also, the matrix elements of any observable that is not explicitly time dependent is also “sta-
tionary” in time for stationary states:
Z
< Ω >E = Ψ∗ (~r, t) Ω Ψ(~r, t) d~r
Z

= ψE (~r) Ω ψE (~r) d~r.

The time derivative operator (for the same Ω) is given by


d d 1
< Ω >E =< “ Ω” >E = < [Ω, H] >E
dt dt ih̄
Z
∗ 1
= ψE (~r) [Ω H − H Ω] ψE (~r) d3 r
ih̄
1
= (E − E ∗ ) < Ω >E =0
ih̄
for real E. So, we indeed have observables independent of time.
Of course, in reality there is no such thing as a stationary state – something must change to
observe the object – such as an electromagnetic transition. Nevertheless, it is useful to construct
this concept as a framework for thinking; just as in classical physics we consider systems in isolation.f
f Just as a frictionless system is a useful abstraction.
CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 76

Stationary States – Momentum Space


Stationary states can also be specified in momentum space. Consider the relation:
~
eik·~r
Z
Ψ(~r, t) = Φ(~k, t) d3 k,
(2π)3/2

where the momentum space stationary state wavefunction is


i
Φ(~k, t) = e− h̄ Et φ(~k).

Here E is constant, independent of k for a stationary state so that we can factor out the time–
dependent phase factor as
~
eik·~r
Z
i
Ψ(~r, t) = e− h̄ Et φ(~k) d~k,
(2π)3/2
and recover
i
= e− h̄ Et ψ(~r).
As before, observables are independent of time:
Z
< Ω >E = Φ∗ (~k, t) Ω Φ(~k, t)d3 k
Z
= φ∗ (~k) Ω φ(~k) d3 k.

The momentum space Schrödinger equation and the associated eigenvalue problem are :

h̄2 2 ~
Z
∂ ~
H Φ(~k, t) = k Φ(k, t) + < ~k | V | ~k 0 > Φ(~k 0 , t)d3 k 0 = ih̄ Φ(k, t),
2m ∂t
and
h̄2 2
Z
k φn (~k) + < ~k | V | ~k 0 > φn (~k 0 ) d~k 0 = En φn (~k)
2m
see H.W. problem.
Another important property is that the expectation value of H is simply E, the energy, which is
real since H is Hermitian. We have already used this property. Proof:
Z Z
∗ ∗
< H >= ψE (~r) H ψE (~r) d3 r = E ψE (~r) ψE (~r) d~r
Z Z
= (H † ψE )∗ ψE d~r = (HψE )∗ ψE d3 r
Z
=E ∗ ∗
ψE (~r) ψE (~r) d~r ... E = E ∗

Note: a non–Hermitian H, such as given by the optical model problem (V1 + iV2 ), would not have a
real energy. For that case, our state is not stationary and we get with E = ER − iEI
i EI
Ψ(~r, t) = e− h̄ ER t e− h̄ t
ψ(~r)
2
| Ψ(~r, t) |2 = e− h̄ EI t | ψ(~r) |2
Z
2
< Ω >E = ψ ∗ (~r) e− h̄ EI t Ω ψ(~r)d~r, etc.

Therefore, the normalization, density, current, expectation values all vanish with time for a EI > 0.
Returning to Hermitian Hamiltonians, we see also that the uncertainty in (∆E) energy vanishes
for stationary states.

(∆E)2 =< (H− < H >E )2 >E =< H 2 >E − < H >2E
CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 77
Z
2 2
= (E − E ) | ψ |2 d3 r = 0

In general: < (H− < H >)n >= 0. Therefore ∆E = 0 and ∆t = ∞. We have a precise energy for
a stationary state, but have no knowledge of the time involved in the motion. Clearly, a particle
localized by forming a wave packet which passes by within a finite ∆t is not a stationary state g
Z
~
H0 Ψ 6= E Ψ for Ψ = ei(k·~r−ωt) φ(~k − ~k0 ) d~k.

3.15 Conditions on ψE (r):


To obtain a solution to the eigenvalue problem for a Hermitian Hamiltonian, boundary conditions
are needed. The differential equation for a local potential case

h̄2 2
− ∇ ψE (~r) + V (r) ψE (~r) = E ψE (~r)
2m
must yield solutions that are physically acceptable, in that the corresponding current, density and
expectation values are well-behaved. The current and density
h̄ ∗ ~
ρE =| ψE (~r) |2 and ~jE = Im (ψE ∇ψE ),
m
have a simple interpretation as a flow of probability –we have ∇ · jE = 0 for stationary states, ρE
doesn’t change. To obtain well behaved current and density (i.e. ρE and jE that are continuous,
single-valued h , nonsingular functions of r, we required that ψE (~r) and ∇ψE (~r) be well behaved,
single-valued continuous and nonsingular everywhere.
Of course, these conditions hold for all wave functions not just stationary ones. Another way of
assuring a reasonable current is seen by taking
h̄ ∗ ∇ψ ∇ψ ∗ h̄ ∇ψ ∇ψ ∗
jE = ψE ψE { − ∗ }= ρE { − ∗ }
2im ψ ψ 2im ψ ψ

and requiring that ψE and ∇ψ


ψE be well behaved, etc.
E

As an example of a discontinuous function, take

ψE = eπiθ(x) ,

where θ is a step function and then


h̄ ~ h̄
jE = ρE {2πi∇θ} = ρE πδ(x) = ∞ at x = 0.
2im m
The current is infinite at one point. Therefore, to assure a finite current we rule out a discontinuous
wave function.
We therefore have the following conditions:
ρ and ~j → finite, continuous, single valued, leads us to require that wave functions are restricted
by ψ and ∇ψ be finite, continuous, and single valued.

3.15.1 Infinite Potentials


There are some cases of particular interest where singular potentials occur for which we need to make
special considerations. The basic procedure is to use the above restrictions on the wave function for
a finite potential and then take the infinite potential limit. For example, an infinite step is obtained
by taking V0 → ∞ in

h̄2 00
− u + V (x)u = Eu
2m
g Wavepackets are therefore superpositions of energy eigenstates.
h In Dirac’s theory of a relativistic particle we shall encounter double-valued states or spinors, see later.
CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 78

For a infinite step potential V (x) = V0 → ∞ for x ≥ 0 and V (x) = 0 for x ≤ 0, we have the solution

u = A sin kx + B cos kx
h̄2 k2
for x < 0 with E = 2m and for x > 0

u = Ce−β|x| + Deβ|x| ,

where s
2m(V0 − E)
β≡ .
h̄2
We have E < V0 since we take the limit as V0 → ∞. Now consider the boundary conditions
• u(∞) → be finite: ... D = 0,
• u(0) continuous: ... B = C,

• u0 (0) be continuous: ... kA = −βC.


Now take V0 → ∞, hence β → ∞. Since we must have u(x) finite for x < 0, we must have
1/2
C → 0 as 1/V0 in order to make −βC → constant. Therefore, we obtain B = C = 0. Finally,
u(x < 0) = A sin kx; at x = 0 we have u(0) = 0, but u0 (0) is free, except that A can be fixed by
normalization conditions. We see in this case that an infinite potential can lead to a discontinuity
in wave functions and their derivatives.
CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 79

3.16 Energy Eigenvalues - Bound State Eigenvalue Problems


To understand the origin of the discrete and later the continuum eigenvalue solutions of our eigen-
value problem
Hψ = Eψ,
let us examine the role played by the boundary conditions on ψ for a fairly general potential.
Consider a one-dimensional local potential V (x) - of finite range.
***** figure inserted here******* In the V (x) = 0 regions, we have positive curvature

u00 2m
= −E 2 > 0,
u h̄
i
for E < 0 solutions. Defining r
2m
β= |E|,
h̄2
the solutions of u00 = β 2 u in the potential free regions are u ∝ e±β|x| . The requirement that u(±∞)
be finite requires us to reject the e+β|x| solution and we have, aside from normalization,

u ∝ e−β|x| .

The tail of the wave function is determined therefore by the negative binding energy.
u00
For points where E = V(x) (called classical turning points) the curvature vanishes u = 0. In
regions where the potential is nonzero and negative:

u00 2m 2m
= (V − E) 2 = (− | V | + | E |) 2 < 0;
u h̄ h̄
that is, the curvature of u is negative or the wave function bends inwards to the axis. For example,
defining a local “constant” r
2m
α = (− | V | + | E |) 2 ,

the solutions behave as oscillatory waves

u ∝ A sin αx + B cos αx.

Now requiring that u = ψ and u0 = ψ 0 be continuous yields only discrete values of E. A procedure
to find the eigenvalues could be to pick a negative E, then integrate in using the Schrödinger equation
- only certain values of E yield proper wave functions. For example, consider a solution √ 2m of the
eigenvalue equation obtained by first selecting the tail of the wave function as u = ψ = +e h̄ |E| |x| .
Then we could use the Schrödinger equation (treated numerically or in some cases analytically) to
find the wave function for smaller x. We integrate in step by step (using difference equation methods
say) to obtain ψ. Suppose our first guess for E yields for a given potential V the following wave
function mismatch: This solution would be rejected since ∇ψ(0) is discontinuous. We could change
the guess for E until a satisfactory solution like
is found for an eigenvalue E1 . Now ψ and ∇ψ are finite, continuous and single valued. Here we
see how the boundary conditions on ψ act to select only discrete eigenvalues for bound ( E < 0
states. q
Another possibility after our first case would be to pick ψ = ± exp(− 2m h̄ | E | | x |) and we
might at first obtain a solution like
*******************************figure to be inserted here
This is clearly rejected since ψ is discontinuous. Therefore, we keep trying by changing E until
a solution with proper ψ and ∇ψ is obtained such as
This process can continue leading to more eigenvalues, more wave functions, and more nodes in
our wave function. The set of bound wave functions ψn found for a given potential is called the set
of bound state eigenfunctions; the corresponding eigenvalues En is called the discrete spectrum.
∂2u
i The notation ψ → u, with u00 denoting the derivative ∂x2
, is often used for 1-D or radial wave functions.
CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 80

x
-10 -5 0 5 10

Figure 3.4: A wave function with discontinuous derivative at x = 0.

x
-10 -5 0 5 10

V
Figure 3.5: An acceptable wave function with continuous derivative and value at x = 0.

-10 -5 0 5 10

Figure 3.6: An acceptable one-node wave function with continuous derivative and value at x = 0.
CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 81

As more nodes develop less binding is obtained and E2 > E1 where ψ2 has more modes than
ψ1 . The nodes imply increase curvature or produce a higher value of ∇2 ψ, or more wave function
oscillations. This is simply the quantum mechanical way of saying more positive kinetic energy
occurs (and hence less binding) as more nodes occur in the eigenfunctions. (We will demonstrate
these remarks about nodes and the usual ordering of energy levels later, along with a discussion of
the number of bound states for a given potential).
The set of bound wave functions have several important properties. First, since they are confined
to a limited region of space their norm || ψn || = finite; in fact, we choose to normalize them as
Z
|| ψn ||≡ | ψn (~r) |2 d3 r = 1 ,

for all bound states n. The above bounded nature of the solutions for negative energy is why we call
such states “bound states.” They are not only normalizeable, but by virtue of H † = H, they are
mutually orthogonal Z
ψn∗ (~r) ψm (~r)d3 r = δnm ,

where δnm = 1 if n = m and δnm = 0 if m 6= n.

3.16.1 Proof of general orthogonality of nondegenerate levels for Hermi-


tian Hamiltonians
We have:
Hψ1 = E1 ψ1
Hψ2 = E2 ψ2
Z Z Z
ψ2∗ Hψ1 d3 r − ψ1 [Hψ2 ]∗ d3 r = (E1 − E2 ) ψ2∗ ψ1 d3 r
Z Z
= ψ2∗ Hψ1 d3 r − (Hψ2 )∗ ψ1 d3 r
Z
= ψ2∗ (H − H † )ψ1 d3 r = 0

Therefore for E1 6= E2 we must have


Z
ψ1∗ ψ2 d3 r = 0,

which demonstrate the orthogonality of nondegenerate states. Degenerate, but distinct solutions ( i.e.
those cases for which say E1 = E2 , but ψ1 6= ψ2 ) can be made orthogonal (Schmidt orthogonalization,
see H.W.). We also choose to normalize states as for example
Z
| ψ1 |2 d3 r = 1
Z
| ψ2 |2 d3 r = 1,

which is generally expressed as


Z
ψn∗ ψm d~r = δnm =< ψn | ψm >≡< n | m >

Another property of bound states is that the wave functions are real functions which gives us zero
current jE = 0, which means that there is no flux escaping.
CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 82

3.17 Stationary States - The Continuum


For the discrete spectrum En < 0 or bound states ψn , the stationary state wave functions are
orthogonal and can be normalized (orthonormal). However, for the stationary states of positive
energy E > 0 any value of E is allowed usually, and hence we call that case the continuum. (There
might be continuum bands, with gaps – see later). The full spectrum consists typically of a discrete
plus continuum spectrum. In some cases only the discrete spectrum occurs (harmonic oscillator), in

Continuous

Gap

Continuous

Bound States

Figure 3.7: A discrete or bound spectrum (E < 0), and a continuous spectrum for (E > 0, ). Gaps
can occur in the continuous spectrum.

other cases only a continuous spectrum occurs (plane wave stationary states, weak potentials with
no bound states). It is also possible in many body problems to have bounded wave function solutions
|| ψE ||= 1 at positive energy E > 0 - called bound states in the continuum. Also the continuum
may be fragmented with gaps - such as occurs in the band theory of solid state physics. We shall
return to these various cases in later problems.
Let me now just emphasize that for the E > 0 spectrum, stationary states are (usually) not
bounded, i.e. Z
k ψE k ≡ | ψE |2 d3 r = ∞ for E > 0.

We recall that one of the failings of the old Bohr-Sommerfeld quantization theory was its inability
to treat non-periodic or continuum problems. Such problems include scattering, resonances, band
theory, etc. Therefore, we must now deal with this difficulty of having unnormalizeable stationary
state solutions. The way to treat these continuum problems is again to invoke the Dirac delta
function and wave packets.

3.17.1 The Continuum - Plane Wave Basis


To understand why continuum E > 0 stationary states are not bounded or normalizeable || ψE ||=
∞, let us consider the simple case of free particles. This discussion will also prepare us for the
general idea of using complete orthonormal basis of states to represent general quantum states and
to the general idea of probability amplitudes. Furthermore, this topic will serve to familiarize us
with the Dirac notation.
The free particle Schroödinger equation

H0 Ψ(~r, t) = ih̄ Ψ(~r, t),
∂t
2

with H0 = − 2m ∇2 , has a general wave packet solution
~
eik·~r − i Ek t ~ ~
Z
Ψ(~r, t) = e h̄ φ(k − k0 ) d~k,
(2π)3/2
CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 83

2

with Ek = 2m k 2 . This, of course, was our beginning point when we decided to use Fourier transforms
to build initial states consistent with the uncertainty principle. It is important to realize that this
wave packet solution is not a stationary state solution. It represents a moving packet with a definite
∆t for passing a point and hence a ∆E 6= 0. Its current and density are not time-independent as we
saw when we studied the spreading and motion of wave packets.
How can we define free particle stationary states? First we need to factor Ψ as
i
Ψ(~r, t) = e− h̄ Et ψE (~r)

with ψE determined by
H0 ψE (~r) = EψE (~r).
h̄2 2
For E > 0, all values of E are permitted by the conditions on ψE ; let us call E = 2m k0 , where k0
can be any real number. Therefore

−∇2 ψE (~r) = k02 ψE (~r)

and therefore
~
eik0 ·~r
ψE (~r) =
(2π)3/2
This plane wave stationary state represents a special case of our general wave packet solution; namely,
the case of taking φ(~k − ~k0 ) localized in k-space

φ(~k − ~k0 ) = δ(~k − ~k0 )

thus
~
i eik0 ·~r
ψ(~r, t) = e− h̄ E(k0 )t ,
(2π)3/2
which is spread out over all space
1
| ψ(~r, t) |2 =| ψE (~r) |2 = ρE (~r) = .
(2π)3
~
The corresponding current is jE = ρE (~r) h̄m
k0
= ρE ~v . Again we have time independent ρE and jE
with ∇ · jE = 0, which is expected for stationary states.
Localized wave packets could be normalized to unity, but our plane wave stationary states are
spread over all space as expressed by the delta function, and we have
~ ~
ei(k0 −k0 )·r
Z Z

k ψE k= ψE ψE d~r = d~r = δ(0) = ∞ !
(2π)3/2

Here we see that continuum E > 0 stationary states are unbounded. They have for plane wave
stationary states the following orthogonality property
Z
ψE∗
r)ψE (~r) d~r = δ(~k00 − ~k0 )
0 (~

The plane wave solutions are also eigenstates of momentum. That property is a natural con-
sequence of having H0 and p~ as commuting and hence compatible observables [~ p, H0 ] = 0. The
eigenvalue equation corresponding to this remark is

p~ ψE (~r) = h̄~k0 ψE (~r);

with E = h̄2 k02 /2m.


CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 84

3.17.2 Momentum Space - Plane Waves Stationary States


We can now express these states in momentum space. The stationary state solution is φ(~k, t) =
i
e− h̄ Et φE (~k).
H0 φE (k) = EφE (k)
or
h̄2 2 h̄2 2
k φE (k) = k φE (k)
2m 2m 0
(k02 − k 2 )φE (k) = 0.
Clearly the solution is φE (k) = δ(~k − ~k0 ) which is an eigenstate of momentum since

p~ δ(~k − ~k0 ) = h̄~k0 δ(~k − ~k0 )

or
p~ φE (~k) = h̄~k0 φE (~k)
Also note that k φE k= δ(k0 − k0 ) = ∞ and φ∗E φE 0 d~k = δ(~k00 − ~k0 )!
R

Nothing new has been said, once it is realized that


~ ~
ik0 ·~
eik·~r r
Z
ψE (~r) = δ(~k − ~k0 ) d3 k = e
(2π)3/2 (2π)3/2

relates our momentum space and configuration space stationary states.


The reason for mentioning this feature of being able to use either the r-space or p-space rep-
resentation is to lead us to the representation - independent notation discussed earlier, i.e., Dirac
notation.

Box Normalization
Another way of treating stationary state plane wave states, which is designed to avoid the difficulties
of the infinite norm, is to use box normalization. We have seen that
Z
|| ψE ||= ∞ or ∗
ψE r = δ(~k0 0 − ~k0 )
0 ψE d~

for plane waves. A finite norm can be defined by restricting our space to be confined to a box
of volume L3 with periodic boundary conditions. The periodic boundary condition ψE (−L/2) =
ψE (L/2) requires that
ei(kx L+ky L+kz L) = 1
or
~kn = 2π ~n
L
n = nx x̂ + ny ŷ + nz ẑ with ni = 0, ±1, ±2 · · · ,
Periodic boundary conditions are used to assure that the operator p~ is Hermitian, since the
surface terms vanish in
Z L/2 Z L/2
∗ h̄ ∂ h̄ ∂ h̄ ∗ L/2
ψE ψE 0 dx = ( ψE )∗ ψE 0 dx + ψE ψE 0 |−L/2
−L/2 i ∂x −L/2 i ∂x i
or ∗
ψE (L/2) ψE 0 (−L/2)
∗ (−L/2) = ψ 0 (L/2) = constant
ψE E

ψE 0 (−L/2) = ψE 0 (L/2)
assures
p† = p !
CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 85

At the same time ψ = 1/L3 and j = ρ~v are indeed stationary since the flux out one wall enters the
opposite wall. The stationary states normalized in this box are thus
~
ψE (~r) = eik·~r /L3/2 = ψn ; H0 ψn = En ψn
with
h̄2 2 h̄2 2π 2 2 2π
En = kn = ( ) n and p~n = h̄ ~n = h̄~kn
2m 2m L L
The spectrum is thus discrete, but becomes continuous as L → ∞
Plane wave states in box normalizations are very useful in many-body studies (see later). For
now we need to observe their property of being a complete orthonormal set in the volume L3 ≡ V
for periodic functions.
The plane waves are normalized and orthogonal when integrating over the finite volume V as
~ ~0
ei(kn −kn )·~r
Z
d~r = δnn0
V V
δnn0 equals 0 for n 6= n0 and 1 for n = n0 .
Any periodic function in the volume V can be expanded in terms of these plane waves as
~
X i
X i eikn ·~r
Ψ(~r, t) = cn e− h̄ En t ψn (~r) = cn e− h̄ En t
n
V 1/2
~
n

Thus ψE form a complete orthonormal basis (ψE ≡ ψn ). (We use the notation Ψ = general function
and ψ = stationary state.) When V → ∞, the above expansion becomes our familiar result of a
wave packet expansion
~
eik·~r
Z
i
Ψ(~r, t) = d~k φ(~k)e− h̄ Ek t
(2π)3/2
where we use the rules
V 1/2 cn → (2π)3/2 φ(k)
and
d~k
Z
1 X

V n (2π)3
Recall that for general φ(k), Ek , the wave packet Ψ is not a stationary state. The last rule for
going to the continuum limit of V → ∞ is based on the observation that we can write cn using the
orthonormality of ψn as: Z
i
cn e− h̄ En t = ψn∗ (~r 0 ) Ψ(~r 0 , t)d~r 0 .
V
Inserting this result into the expansion of Ψ, one finds
Z X
Ψ(~r, t) = ( ψn (~r)ψn∗ (~r 0 ))d~r 0 Ψ(~r 0 , t),
V n

from which it follows that X


ψn (~r) ψn∗ (~r 0 ) = δV (~r − ~r 0 )
n
behaves as a δ-function for the finite volume. As V → ∞ this result becomes
1 X i~kn ·(~r−~r 0 )
lim δV (~r − ~r 0 ) = δ(~r − ~r 0 ) = lim e
V →∞ V →∞ V
n
~ 0
eik·(~r−~r )
Z
= d~k
(2π)3
with the substitution
d~k
Z
1 X

V n (2π)3
being used to go to the continuum limit.
Having discussed this important example of a continuum and a complete orthonormal set, let us
now generalize the discussion to the case of a full Hamiltonian H = H0 + V.
CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 86

3.18 Complete Orthonormal Basis


The stationary state solutions of the Schrödinger equation provide a complete orthonormal basis of
states for a Hermitian Hamiltonian H = H0 + V. This fact is an essential part of quantum theory as
i
we shall see. The stationary states ΨE (~r , t) = e− h̄ Et ψE (~r) are, for a local potential, solutions to

h̄2 2
[− ∇ + V (r) ] ψE (~r) = E ψE (~r),
2m
with proper boundary conditions on ψE yielding a discrete k ψE k= 1 and a continuous k ψE k= ∞
spectrum. The above problem is a Sturm-Liouville equation, which can be shown to yield a complete
set of states. Instead of proving that theorem, let us accept it. Later, we shall see that completeness
(when applied to spin and other quantum numbers) becomes a physical not just a mathematical
principle.
Using completeness, any general function Ψ(~r, t) can be expanded in terms of the discrete and
continuum solutions as
Z
i i
X
Ψ(~r, t) = cn e− h̄ En t ψn (~r) + dλ c(λ) e− h̄ Eλ t ψλ (~r).
n

Here n denotes the quantum numbers labeling the states; for now n can be thought of as the
energy alone, but we will supplement this with additional quantum numbers later. The integral over
dλ denotes integration over an appropriate continuum variable, such as the momentum or energy
variables.
We have Ψ as a solution to the Schrödinger equation since using Hψn = En ψn
Z
∂ X i i
(H − ih̄ Ψ) = (En − En ) cn e− h̄ En t ψn + dλ (Eλ − Eλ ) cλ e− h̄ Eλ t ψλ = 0.
∂t n

The orthonormal property of these stationary states is for discrete bound states
Z
ψn∗ (~r) ψn0 (~r) d~r = δnn0 ,

for continuum states Z


ψλ∗ (~r)ψλ0 (~r) d~r = δ(λ − λ0 ),

and Z
ψλ∗ (~r ) ψn (~r ) d~r = 0

Using these properties, we can express the coefficients cn as


Z
i
cn e− h̄ En t = ψn∗ (~r 0 )Ψ(~r 0 , t) d~r 0

and Z
i
cλ e− h̄ Eλ t = ψλ∗ (~r 0 )Ψ(~r 0 , t) d~r 0 .

These coefficients are called probability amplitudes for reasons to be discussed soon. Inserting these
expressions into the completeness expansions
Z X Z
Ψ(~r, t) = d~r 0 { ψn (~r ) ψn∗ (~r 0 ) + dλ ψλ (~r ) ψλ∗ (~r 0 ) } Ψ(~r 0 , t),
n

leads us to identify the closure relationship


X Z
ψn (~r ) ψn∗ (~r 0 ) + dλ ψλ (~r ) ψλ∗ (~r 0 ) = δ(~r − ~r 0 ).
n

The above closure property is the same property deduced earlier for the plane wave basis except
for the inclusion of the bound states (in the discrete sum term). For convenience, we will often use
CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 87

the summation to represent a sum over the discrete (bound state) indices plus an integration over
the continuum:
X Z X
+ dλ → .
n n

Formally, closure will therefore be written as


X
| n >< n |= 1.
n

Note that closure follows from orthonormalty plus completeness of a basis.

Dirac Notation – Revisited


The continuum states have been treated using the Dirac delta function. We have discussed the
completeness and closure properties of stationary states using the orthonormality properties based
on the infinite, delta function normalization for continuum states.
Now let me briefly discuss these ideas using Dirac notation. The main advantages of this notation
are its brevity and its emphasis on describing quantum mechanics in a representation independent
manner (i.e. valid in both r and p space). It is the language that we can generalize to spin and
other quantum concepts. The basic idea is to represent functions as vectors in an abstract Hilbert
space. The components of the vectors define functions of ~r or p~.

< ~r | Ψ(t) >≡ Ψ(~r, t)

and Z
| Ψ(t) >= | ~r > d~r < ~r | Ψ(t) >

The dual or adjoint space yields


< Ψ(t) | ~r >≡ Ψ∗ (~r, t).
Projection to obtain components in p-space yields:

< ~k | Ψ(t) >≡ Φ(~k, t)

The dual or adjoint space yields


< Ψ(t) | ~k >≡ Φ∗ (~k, t)
and Z
| Ψ(t) >= < Ψ(t) | ~k >< ~k | d~k .

The best way to learn how to deal with these abstract Hilbert space notations is simply to see
examples of how to write our familiar equations in this notation. For example, the Schrödinger
equation is in abstract form

H | Ψ(t) >= ih̄ | Ψ(t) > .
∂t
Projection onto < ~r | returns us to function space and

h̄2 2 ∂
< ~r | H | Ψ(t) >= [− ∇ + V (r)]Ψ(~r, t) = ih̄ Ψ(~r, t),
2m ∂t
where the steps Z
< ~r | H | Ψ(t) >= < ~r | H | ~r 0 > d3 r0 < ~r 0 | Ψ(t) >

and
h̄2 2
< ~r | H | ~r 0 >= [− ∇ + V (r)] δ(~r − ~r 0 )
2m r
have been used.
The Schrödinger equation can also be written as
i
| Ψ(t) >= e− h̄ Ht | Ψ(0) > .
CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 88

For stationary states, using Dirac notation we have


i
| Ψ(t) >= e− h̄ En t | ψn >,
which corresponds to:
i i
< ~r | Ψ(t) >= Ψ(~r, t) = e− h̄ En t ψn (~r) = e− h̄ En t < ~r | ψn >
These stationary states satisfy the time–independent Schrödinger equation
H | ψn >= En | ψn > or H | n >= En | n >
The advantage of the above notation is that the p-representation case is also included since
i i
< ~k | Ψ(t) >= Φ(~k, t) = e− h̄ En t φn (~k) = e− h̄ En t < ~k | ψn > .
Instead of writing | ψn >, we can for convenience label the stationary states by | n >≡| ψn > .
Using that convention, the orthonormality property is expressed simply as
< n | n0 >= δnn0 ,
where δnn0 is a Kronecker delta function for discrete states, but a Dirac delta function for continuum
states. Note that Z
< n | n0 >= < n | ~r > d~r < ~r | n0 >
Z
= d~r ψn∗ (~r)ψn0 (~r) = δnn0 ,

We use the property discussed earlier that


Z
d~r | ~r >< ~r |= I.

The closure property is expressed formally as


X
| n >< n |= I,
n

which includes a sum over discrete and an integration over continuum states. The explicit form of
this relation is X
< ~r | n >< n | ~r 0 >=< ~r | ~r 0 >= δ(~r − ~r 0 )
n
X
= ψn (~r) ψn∗ (~r 0 ),
n
as before.
The completeness relation can also be written formally as
i
X X
| Ψ(t) >= | n >< n | Ψ(t) >= | n > e− h̄ En t cn .
n n

In the r-representation this is


i
X
< ~r | Ψ(t) >= Ψ(~r, t) = ψn (~r)e− h̄ En t cn .
n

The coefficient cn is again given by


i i i i
X
cn e− h̄ En t =< n | e− h̄ Ht | Ψ(0) >= < n | e− h̄ Ht | n0 >< n0 | Ψ(0) >= e− h̄ En t < n | Ψ(0) > .
n0

− h̄i Ht − h̄i En t
Thus, < n | e | n0 >= e δnn0 and
Z
cn =< n | Ψ(0) >= ψn∗ (~r)Ψ(~r, 0)d3 r.

The above relation tells us that the probability amplitude for finding the system in an eigenstate n
is given by the overlap of ψn with the initial state Ψ(~r, 0). Now we discuss, using Dirac notation,
why it is appropriate to call cn a probability amplitude.
CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 89

3.18.1 Probability Amplitudes


For a general solution of the time–dependent Schrödinger equation, Ψa (~r, t), which represents a
general state of the system, the expectation value of the Hamiltonian yields the average energy
Z
Eav =< H >a = Ψ∗a (~r, t)HΨa (~r, t) d3 r

X
=< Ψa | H | Ψa >= < ψa | n >< n | H | n0 >< n0 | Ψa >
n,n0
X X
= |< ψa (t) | n >|2 En = | cna |2 En .
n n

We have used H | n0 >= En0 | n0 > and hence with < n | n0 >= δnn0 , we have < n | H | n0 >=
i
En δnn0 . Also used are the properties: < Ψa (t) | n >=< n | Ψa (t) >∗ , and < n | Ψa (t) >= e− h̄ En t cn ,
and cn =< n | Ψa (0) > . These steps have been presented in Dirac notation, but the explicit integrals
can also be given.
The physical interpretation of the above result is that in a given measurement the energy
eigenvalue En is observed, for a general state of the the system Ψa , with a probability | cna |2 ,
i.e.| cna |2 gives the probability of measuring the eigenvalue En in the state Ψn . Therefore, we call
cna =< n | Ψa (t) > the probability amplitude. Note that

|< n | Ψa (t) >|2 =|< n | Ψa (0) >|2

is independent of time since once the state Ψa is prepared, and allowed to evolve without disturbance,
the extent to which it contains a given eigenstate remains unchanged. ( Here we deal with eigenstates
of the full Hamiltonian. Later, when we introduce perturbation theory, we will have transitions,
induced by perturbing interactions, between eigenstates of the unperturbed Hamiltonian.)
These remarks can be generalized to apply to any other observable for which an eigenvalue
equation exists (such as angular momentum and other constants of the motion) of the form

Ω |>= ωn | n >

Then X X
< Ω >a =< Ψa | Ω | Ψa >= | cn |2 ωn = |< n | Ψa >|2 ωn
n n
2
with | cn | again giving the probability of measuring the eigenvalue ωn in the general state Ψa .
We shall return to this topic again. It should be noted now that one always observes the
eigenvalues, with a probability given by |< a | n >|2 =|< Ψa | n >|2 . Hence, the eigenvalue equation
plays a fundamental role in our understanding of the measurement process.

3.18.2 Projective measurement and Wavefunction Collapse


to be inserted
CHAPTER 3. SCHRÖDINGER EQUATION–FREE PARTICLE MOTION 90

3.18.3 Matrix Mechanics


The general state of a quantum mechanical system is specified by giving the wave function Ψa (~r, t),
which is a solution of the Schrödinger equation ih̄Ψ̇ = HΨ. The average value of observables ob-
tained in repeated measurements on systems described by Ψa is then obtained by evaluation of the
expectation value < Ψa | Ω | Ψa > . Each measurement in fact yields one of the eigenvalues ωn
(or En ). Repeated measurements then indicate the probabilities for finding a particular eigenvalue
En in the state Ψa , i.e. |< n | Ψa >|2 . The prediction of such results requires the evaluation of
< n | Ψa > or of the eigenfunction | n > and the corresponding eigenvalues En . The preparation (and
boundary conditions) of the system in an experiment dictates what Ψa is. One way of determining
the eigenvalues and eigenfunctions is to solve

H | Ψn >= En | Ψn >,

subject to boundary conditions on Ψn (r) as a differential equation.


Another approach is to use a matrix formulation. Suppose, the eigenstates and eigenvalues of H
exist but we do not know them. We know they form a complete orthonormal set. The eigenvalues
En and the probability amplitudes cn will be found later to be described as a matrix problem. The
eigenvalues En and eigenfunctions | n > could then be solved by solving a matrix problem. For
example, one gets the usual condition that det (H - E 1) = 0. Both En and cn ’s can be determined.
(Note only the energy eigenvalues and the probability amplitude are involved once the Hamiltonian
matrix is specified.) Examples of the use of matrix methods will be given later.
Also I would like to return later to a discussion of the Historical development of matrix mechanics,
including an evaluation of when it is more convenient to use one approach or the other. The present
discussion is dedicated only to showing that the existence of a complete orthonormal set provides
the means of relating the wave and matrix mechanics approaches.
Conclude:
A complete orthonormal set of eigenstates led us to the “Expansion Postulate of Quantum Me-
chanics” and to the matrix formulation.
The expansion postulate (as stated nicely by Merzbacher, p. 153) is
P by a Hermitian operator with eigenfunctions ψ1 , · · · · · · ψn
“Every physical quantity can be represented
and every physical state Ψ by a sum Ψ = n cn ψn , where
Z
cn = ψn∗ Ψ dr;

| cn |2 is the probability of finding the value An for an observable A in the state Ψ.


Chapter 4

Motion in Three Dimensions

Instead of developing further the formalism of quantum mechanics, I will instead discuss applications
of the Schrödinger equation to various problems of general interest. Some applications, especially
to one-dimensional problems, will be given in the problems. The development of matrix mechanics
and transformation theory will be returned to later.

4.1 The Two-Body Problem


The motion of two interacting bodies of masses m1 and m2 is described by the Hamiltonian

p21 p2
H = H0 + V (| r~1 − r~2 |) = + 2 + V (r) = H1 + H2 + V.
2m1 2m2
The relative coordinate is defined by ~r = ~r1 − ~r2 = ~r12 . In the r-space representation, one must solve
the differential equation in 6-dimensions (3N).

h̄2 2 h̄2 2 ∂
[− ∇1 − ∇2 + V (r12 ) ] Ψ(~r1 , ~r2 , t) = ih̄ Ψ(~r1 , ~r2 , t) .
2m1 2m2 ∂t
For N ≥ 2 bodies, the natural generalization is
N
X h̄2 2 X ∂
[ (− ∇i ) + V (rij ) ] Ψ(~r1 · · · ~rn , t) = ih̄ Ψ(~r1 · · · , ~rn , t) .
i
2mi ∂t
i6=j

Here it is clear that Ψ is a function in the 3N-dimensional configuration space (not in ordinary 3-D
space). It is one of the significant aspects of the “new” quantum theory that approximate solutions
to the N ≥ 2 problems can be found. Also note that a single time for specifying the state of all N
bodies is used as part of our nonrelativistic description.
We can introduce the momentum space version of the two-body problem using Fourier transforms
Z
1 ~ ~
Ψ(~r1 , ~r2 , t) = ei(k1 ·~r1 +k2 ·~r2 ) Φ(~k1 , ~k2 , t) d3 k1 d3 k2 ,
(2π)3
or in Dirac notation
Z
< ~r1 , ~r2 | Ψ(t) >= d3 k1 d3 k2 < ~r1 , ~r2 | ~k1 , ~k2 >< ~k1 , ~k2 | Ψ(t) >,

where | ~k1 ~k2 > can really be split as | ~k1 >| ~k2 >, i.e. as a product of plane waves.
The Schrödinger equation in momentum space is therefore:

p1 2 p2 2
Z
0 0 0 0
( + )Φ(k1 , k2 , t) + d3 k1 d3 k2 < k~1 , k~2 | V | k~1 k~2 > Φ(k~1 , k~2 , t)
~ ~
2m1 2m2

= ih̄ Φ(k~1 , k~2 , t),
∂t
CHAPTER 4. MOTION IN THREE DIMENSIONS 92

with p~1 = h̄~k1 , p~2 = h̄~k2 and


Z
1 ~ ~ 1 ~0 ~0
< ~k1 , ~k2 | V | ~k10 ~k20 >= e−i(k1 ·~r1 +k2 ·~r2 ) V (r12 )d3 r1 d3 r2 ei(k1 ·~r1 +k2 ·~r2 ) .
(2π)3 (2π)3

Six–dimensional differential (or integral) equations are rather difficult to solve unless we reduce the
dimensionality using the physical ideas of center of mass and spherical symmetry.

4.2 Center of Mass


Let us consider the center of mass for two noninteracting particles, for which H = H1 + H2 =
p21 p22
2m1 + 2m2 . In momentum space we have

p21 p2 ∂
( + 2 ) Φ(~k1 , ~k2 , t) = ih̄ Φ(k~1 , k~2 , t),
2m1 2m2 ∂t
or in ~r−space
h̄2 2 h̄2 2 ∂
[− ∇1 − ∇ ] Ψ(~r1 , ~r2 , t) = ih̄ Ψ(~r1 , ~r2 , t).
2m1 2m2 2 ∂t
For two noninteracting bodies, it is clear that one has separate equations for each body H1,2 Ψ1,2 =
ih̄Ψ̇1,2 , and hence the wave functions are simple products Ψ = Ψ1 (~r1 , t) Ψ2 (~r2 , t) and


(H1 + H2 )Ψ1 Ψ2 = ih̄(Ψ̇1 Ψ2 + Ψ1 Ψ̇2 ) = ih̄ Ψ1 Ψ2 = ih̄Ψ̇.
∂t
For average energies, the expectation values give a sum

Eav =< Ψ | H1 + H2 | Ψ >=< Ψ1 | H1 | Ψ1 >< Ψ2 | Ψ2 >

+ < Ψ2 | H2 | Ψ2 >< Ψ1 | Ψ1 >= E1 + E2 .


The above steps show that for energy eigenstates (H1,2 ψ1,2 = E1,2 ψ1,2 ) the energy of the above two
noninteracting systems is simply the sum E = E1 + E2 , as expected.
For N noninteracting particles the corresponding result is
N
X
H= Hi
1

with
Ψ = ΠN
1 Ψi (ri , t)

(a product wave function) and


N
X
E= Ei .
1
a

Isolated, or noninteracting, systems are therefore represented by product wave functions (also
called separable wave functions.)
Product wave functions in which the center of mass motion is separated out from the internal
motion can be obtained by first defining relative and center of mass momenta and coordinates. In
the momentum representation, the momenta are defined in the usual way (as numbers).

m2 p~1 − m1 p~2
p~12 = p~ = h̄~k = → µ(~v1 − ~v2 )
M
and
P~ = h̄K
~ = p~1 + p~2 .
a The symmetry condition of many–body wavefunctions under particle exchange for identical particles will be

discussed later.
CHAPTER 4. MOTION IN THREE DIMENSIONS 93

with total M = m1 + m2 and reduced µ = m1 m2 /M masses. It follows that


m1 ~ m2 ~
p~1 = P + p~ ; p~2 = P − p~
M M
and
p1 2 p2 2 P2 p2
+ = + .
2m1 2m2 2M 2µ
Introducing the relative and c.m. coordinates

~r12 = ~r = ~r1 − ~r2

~ = m1~r1 + m2~r2
R
M
~+ m2 ~ − m1 ~r
~r1 = R ~r ; ~r2 = R
M M
one finds that
p~1 · r~1 + p~2 · r~2 = P~ · R
~ + p~ · ~r.

With unit Jacobians: d~ p2 = dP~ d~


p1 d~ ~
p and d~r1 d~r2 = d~r dR.
The noninteracting particles can therefore be represented by a separable wave function since

P2 p2
H = H1 + H2 = + = Hc.m. + Hint ,
2M 2µ

~ t)Φ(~k, t),
and Ψ = χcm (K,

~ t)Φ(~k, t) = ih̄ ∂ χcm (K,


(Hcm + Hint )χcm (K, ~ t)Φ(~k, t),
∂t
with Hcm χcm = ih̄∂χcm /∂t and Hint Φ = ih̄∂Φ/∂t . Clearly the average energy adds as Eav =
cm int
Eav + Eav . The motion has been separated into c.m. and internal (int) or relative motion.
So far, no operators have appeared, since we worked in momentum space. Now we can introduce
~r-space and find
Z
1 i(K· ~ ~
~ R+k·~
r) ~ t) Φ(~k, t) d3 K d3 k = χcm (R,
~ t) Φ(~r, t),
Ψ(~r1 , ~r2 , t) = e χcm (K,
2π)3

with
h̄2 2 h̄2 2 h̄2 2 h̄2 2
H=− ∇1 − ∇2 = − ∇R − ∇ .
2m1 2m2 2M 2µ r
Now the interaction can be immediately introduced
For later use it is worth noting that in momentum space

< ~k1~k2 | V | ~k10 ~k20 >= δ(K ~ 0 ) < ~k | V | ~k 0 >


~ −K

where for a local potential


~ ~0
e−ik·~r eik ·~r
Z
< ~k | V | ~k 0 >= d3 r V (r) .
(2π)3/2 (2π)3/2

Again the c.m. motion factors out and we have

P 2 p2
Z
[ ~ ~
+ ]χ(K, t)Φ(k, t)+ δ(K− ~ K ~ 0 , t) Φ(~k 0 , t) = ih̄ ∂ χ(K,
~ 0 ) < ~k | V | ~k 0 > d3 k 0 d3 K 0 χ(K ~ t)Φ(~k, t)
2M 2µ ∂t
which separates into the c.m motion:

P2 ~ t) = ih̄ ∂ χ(K,
~ t)
χ(K,
2M ∂t
CHAPTER 4. MOTION IN THREE DIMENSIONS 94

and the relative motion:


p2
Z

Φ(~k, t) + < ~k | V | ~k 0 > d3 k 0 Φ(~k 0 , t) = ih̄ Φ(~k, t).
2µ ∂t
These equations in momentum space will be very useful in our later discussion of scattering.
The separation given above for the potential can also be given as
~ −R
< ~r1~r2 | V | ~r1 0~r2 0 >= δ(R ~ 0 ) < ~r | V | ~r 0 >;

for a local potential we have < ~r | V | ~r 0 >= δ(~r − ~r 0 )V (r). Again the c.m. motion is factored out.
The stationary state for the motion of the c.m. is obviously
~
iK0 ·R ~
~ t) = e− h̄i E(K0 )t e
Ψ(R, ,
(2π)3/2

with E(K0 ) = h̄2 K02 /2M, and

~ t) = e− h̄i E(k0 )t δ(K


χc.m. (K, ~ −K
~ 0 ),

in momentum space.
CHAPTER 4. MOTION IN THREE DIMENSIONS 95

4.3 Orbital Angular Momentum


The angular momentum operator for each particle
~
L(1) = ~r(1) × p~(1) ~
L(2) = ~r(2) × p~(2)

can be added to define the total orbital angular momentum of the two-body system
~
L = L(1) ~
+ L(2).

With the operators p~(1) and p~(2) brought inside the Fourier transform integrals (and ~r(1) and ~r(2)
kept outside), the classical result follows immediately as

~ =R
L ~ × P~ + ~r × p~ = L
~ c.m + L
~ int .

In these steps we assume that the two particles are physically distinct and hence

[p(1), p(2)] = 0 [L(1), L(2)] = 0.

Additional Remarks about Center of Mass Problems


We could consider the density and current rules for relative motion ρ = ρcm ρint and j = jc.m. + jint .
Also, the case
1 e1 ~ 2 1 e2 ~ 2
H= p1 − A
(~ 1) p2 − A
(~ 2 ) + e 1 Φ 1 + e 2 Φ2 ,
2m1 c 2m2 c
for external fields will be discussed later.
The separation of the c.m. in relativistic cases requires further discussion (see Foldy’s papers,
1971).
The Einstein-Podolski-Rosen paper discusses the basic interpretation of q.m. with an example
(among others) of

[~
p1,2 , ~r1,2 ] = ⇒
i
and associated uncertainty principles.
CHAPTER 4. MOTION IN THREE DIMENSIONS 96

4.4 Central Potentials


The Schrödinger equation has been reduced to a 3-D differential equation (in momentum space it is
a 3-D integral equation)
h̄2
[− ∇2 + V (r)]Ψ(~r, t) = ih̄Ψ̇(~r, t) . (4.1)

The potential is assumed to be central and local V (| ~r |). (Nonlocal and noncentral potentials will
be discussed later). For central potentials, the wave function can be further separated into a radial
times an angular wave function.

u(r)
ψE (~r) = R(r)Y (r̂) = R(r)Y (θ, φ) = Y (r̂), (4.2)
r
i
where we have introduced the energy eigenstates Ψ(~r, t) → e− h̄ Et ψE (~r ). To see why it is possible
to make the above separation, let us recall thatb

1 ∂ 2 ∂ 1 1 ∂ ∂ 1 ∂2
∇2 = r + ( sin θ + ) (4.3)
r2 ∂r ∂r r2 sin θ ∂θ ∂θ sin2 θ ∂φ2
in spherical coordinates.
x = r sin θ cos φ
y = r sin θ sin φ
z = r cos θ
r2 = x2 + y 2 + z 2
y
tan φ =
x
z
cos θ =
r
d~r = r2 drdΩ
dS
dΩ = sin θdθ dφ = . (4.4)
r2
We can define (in obvious preparation for a discussion of orbital angular momentum) the following
operators:
h̄2 ∂ ∂ ∂2
L2 = − 2 [sin θ ∂θ sin θ ∂θ + ∂φ2 ]
sin θ
h̄ 1 ∂ h̄ ∂ 1
pr = r= ( + ). (4.5)
i r ∂r i ∂r r
Recall that pr is Hermitian provided the surface terms vanish; however, pr is not an observable
since it satisfies no eigenvalue equation (See Messiah). The operator p2r can be written in several
useful forms
1 ∂ 1 ∂ 1 ∂2
p2r = −h̄2 r r = −h̄2 r
r ∂r r ∂r r ∂r2
1 ∂ ∂ ∂2 2 ∂
= −h̄2 (1 + r ) = −h̄2 ( 2 + )
r ∂r ∂r ∂r r ∂r
1 ∂ ∂
= −h̄2 2 r2 . (4.6)
r ∂r ∂r
Using p2r and L2 , we can write ∇2 as
L2
−h̄2 ∇2 = p2r +
r2
b In the classical Hamiltonian...., we can not simply use the following replacements:
h̄ ∂ h ∂ h̄ ∂
pr = pθ = , pφ =
i ∂r i ∂θ i ∂φ
in above classical Hamiltonian. Need to make H hermitian and watch order of operators. Use of the correct curvilinear
form for ∇2 leads to correct form! (See Pauli Handbuck der Physik on quantization in noncartesian coordinates).
CHAPTER 4. MOTION IN THREE DIMENSIONS 97

and:
h̄2 2 p2 L2
H=− ∇ + V (r) = [ r + V (r) + ] (4.7)
2µ 2µ 2µr2
The operators separate, with L2 acting on θ, φ only and hence

p2r L2
HR(r)Y (θ, φ) = [ + V (r) + ]R(r)Y (θ, φ), (4.8)
2µ 2µr2
and
r2 p2r 1 1 2 λ
−r2 E + [ + V (r)]R = − L Y =− = constant! (4.9)
R 2µ 2µ Y 2µ
The last step follows from the structure of the above equation; namely, that if a function of r equals
a function of θ and φ for all values of these variables, it must be equal to a constant, which we call
λ
− 2µ . We deduce the radial equation

p2r λ
[ + V (r) + ]R(r) = ER(r), (4.10)
2µ 2µr2
and the angular equation
L2 Y (θ, φ) = λY (θ, φ). (4.11)
Now Y (θ, φ) can also be separated Y (θφ) = P (θ)Φ(φ)

h̄2 ∂ ∂ ∂2
− 2 [sin θ ∂θ sin θ ∂θ + ∂φ2 ]P (θ)Φ(φ) = λP (θ)Φ(φ)
sin θ
h̄2 ∂ ∂ h̄2 ∂ 2
− [sin θ sin θ ]P (θ) − sin2 θ λ = Φ(φ) = constant ≡ −h̄2 m2` . (4.12)
P (θ) ∂θ ∂θ Φ(φ) ∂φ2

The solution for Φ is Φ = eim` φ / 2π, with m` = 0, ±1, ±2, chosen to make Φ single- valued. c
The equation for P (θ) can be rearranged and shown to be Legendre’s equation

∂ ∂ λ
[sin θ sin θ + sin2 θ]P (θ) = 0, (4.13)
∂θ ∂θ h̄2

with ξ ≡ cos θ, and hence ∂


∂ξ = − sin1 θ ∂θ

,we set λ ≡ h̄2 `(` + 1), and find various forms for Legendre’s
equation:
∂ λ
[(1 − ξ 2 ) (1 − ξ 2 ) + 2 (1 − ξ 2 ) − m2` ]P (ξ) = 0
∂ξ h̄
2
2 ∂ ∂ m2`
[(1 − ξ ) 2 − 2ξ + `(` + 1) − ]P (ξ) = 0
∂ξ ∂ξ 1 − ξ2
∂ ∂ m2`
[ (1 − ξ 2 ) − + `(` + 1)]P (ξ) = 0
∂ξ ∂ξ 1 − ξ2

Properties of the Legendre functions P (ξ) are give in Abramowitz and Stegun and in many other
sources. Also see HW problems. There are regular Legendre P`m` (z) and irregular Legendre Q`m` (z)
functions. We require that the wave function be finite over the full range of angles, so we reject the
irregular solutions. d For the case of integer ` and m` these function are polynomial functions of
order ` − |m` |. Indeed the integral
Z 1
dxP`m` (x)P`m` (x),
−1

where x = cos θ, is finite provided ` is a positive integer and m` = 0, ±1, ±2 · · · .


c See later for a better reason for having integer m values for orbital angular momentum. The upper limit for
`
| m` | is ` as will be shown later.
d These Q
`m` (z) functions will reappear later when we consider the partial wave decomposition of Yukawa poten-
tials.
CHAPTER 4. MOTION IN THREE DIMENSIONS 98

The Radial Equation


The radial stationary state Schrödinger equation can be written in several standard forms:

p2r h̄2 `(` + 1)


[ + V (r) + ]Rn` (r) = En` Rn` (r)
2µ 2µ r2
1 ∂2 `(` + 1) 2µ un` 2µ un`
{− r+[ + 2 V (r)]} = En`
r ∂r2 r2 h̄ r h̄2 r
∂2 `(` + 1) 2µ
− 2 un` + [ + 2 V (r)]un` = En` un` (4.14)
∂r r2 h̄
Note that the energy for a central potential is independent of m` . It is convenient to define the radial
function un` (r) = rRn` (r), which is subject to the boundary condition un` (0) = 0, since Rn` (0) is
finite and for bound states Z ∞
un` (r)un` (r)dr = 1.
0
Note that for bound states
Z ∞ Z
2 2 ∗
kψk = drr Rn` (r) dΩ Y`m Y
` `m`
0


R∞
drr2 Rn`
2
R
has a bounded dΩ Y`m Y
` `m`
and thus kψk is bounded provided 0
(r) is finite.

4.5 Orbital Angular Momentum


The central force problem has led us to consider eigenstates of the operator

h̄2 ∂ ∂ ∂2
L2 = − 2 [ sin θ sin θ + ]
sin θ ∂θ ∂θ ∂φ2
with the eigenvalue equation

L2 Y`m` (θ, φ) = h̄2 `(` + 1)Y`m` (θ, φ).

We now identify L~ = ~r × p~ as the orbital angular momentum operator for the relative motion of two
~
bodies. Recall L =⇒ R ~ × P~ + ~r × p~ = L
~ c.m. + L
~ int .
Let me briefly review some of the important properties of this operator. We have
~ = ~r × p~ = x̂Lx + ŷLy + ẑLz ,
L

and in determinant form


x̂ ŷ ẑ
~

L = x y z .
∂ ∂ ∂


∂x ∂y ∂z

It is readily shown that


[Lx , Ly ] = ih̄Lz [Ly , Lz ] = ih̄Lx
and
[Lz , Lx ] = ih̄Ly ,
~ = Lx
which we can write as an operator equation ih̄Li = ijk Lj Lk or as ih̄L ~ L.
~ (For example,
ih̄Lz = Lx Ly − Ly Lx ).
Proof:
h̄ ∂ ∂ ∂ ∂ h̄ ∂ ∂ ∂ ∂
[Lx , Ly ] = ( )2 [ y − z ,z −x ] = ( )2 ([ y , z ] + [ z ,x ]) =
i ∂z ∂y ∂x ∂z i ∂z ∂x ∂y ∂z
h̄ h̄ ∂ ∂
[y −x ] = h̄iLz ,
i i ∂x ∂y
and similar steps for [Lz , Lx ]. It is simple to prove all of these using the ijk symbol (see HW).
CHAPTER 4. MOTION IN THREE DIMENSIONS 99

As further consequences of our basic commutator [p, r] = h̄/i, we have with L2 = L2x + L2y + L2z

[L2 , Lx ] = [L2y + L2z , Lx ] = Ly [Ly , Lx ] + [Ly , Lx ]Ly + Lz [Lz , Lx ] + [Lz , Lx ]Lz

= −h̄iLy Lz − h̄iLz Ly + ih̄Lz Ly + ih̄Ly Lz = 0


Also [L2 , Ly ] = [L2 , Lz ] = 0 or
~ =0
[L2 , L] [L2 , L± ] = 0,
where
L± ≡ Lx ± iLy .
are Raising and Lowering operators.
We have L2 and Lz as mutually compatible observables; only one component can be observed
precisely because of [Lx , Ly ] = ih̄Lz etc. The uncertainty principle corresponding to [L2 , Lz ] = 0 is
∆L2 ∆Lz = 0 so both L2 and Lz can be precisely and simultaneously determined and hence used to
characterize the state of the system.
In the above step, we used L2 ≡ L2x + L2y + L2z , we must now show that this definition is the
same as the earlier one

h̄2 ∂ ∂ ∂2
L2 = − 2 2 2
2 [sin θ ∂θ sin θ ∂θ + ∂φ2 ] = Lx + Ly + Lz .
sin θ

~
Explicit forms for L
The attached appendix contains a proof of the following
h̄ ∂
Lz =
i ∂φ
∂ ∂
Lx = ih̄(sin φ + cos φ cot θ )
∂θ ∂φ
∂ ∂
Ly = ih̄(sin φ cot θ − cos φ )
∂φ ∂θ
∂ ∂
L± = ±h̄e±iφ { ± i cot θ }
∂θ ∂φ
1 ∂ ∂ 1 ∂2
L2 = −h̄2 { sin θ + }
sin θ ∂θ ∂θ sin2 θ ∂φ2
The relation for Lz shows that its eigenvalue equation holds as

Lz Φm = mh̄Φm

Φm = eimφ / 2π
or
< Φm0 | Lz | Φm >= δmm0 h̄m.
< Lz >= mh̄ and ... ∆Lz = < L2z > − < Lz >2 = 0.
p

Therefore, the average value of < Lz > is precisely Lz . Said another way Lz is a constant of the
motion since for a central potential
dLz
= ih̄[Lz , H] = 0
dt
d
... < Lz >= 0 or < Lz >= constant = m` h̄.
dt
CHAPTER 4. MOTION IN THREE DIMENSIONS 100

Appendix - Explicit Forms for Lx , Ly , Lz and L2


x = r sin θ cos φ
y = r sin θ sin φ
z = r cos θ
p
r = x2 + y 2 + z 2
tan φ = y/x
cos θ = z/r

h̄ ∂ ∂
Lx = (y −z )
i ∂z ∂y
h̄ ∂ ∂
Ly = (z −x )
i ∂x ∂z
h̄ ∂ ∂
Lz = (x −y )
i ∂y ∂x
Use of chain rule:
∂ ∂r ∂ ∂θ ∂ ∂φ ∂
|yz = |yz + |yz + |yz etc.
∂x ∂x ∂r ∂x ∂θ ∂x ∂φ

∂r x
= = sin θ cos φ
∂x r
∂r y
= = sin θ sin φ
∂y r
∂r z
= = cos θ
∂z r
∂θ zx 1
= = cos θ cos φ
∂x sin θr3 r
∂θ zy 1
= = cos θ sin φ
∂y sin θr3 r
∂θ z2 1 1 1
= − = − sin θ
∂z sin θ r3 r sin θ r
∂φ y r sin θ sin φ sin φ
= − cos2 φ 2 = − cos2 φ 2 2 =−
∂x x r sin θ cos2 φ r sin θ
∂φ 1 cos φ
= cos2 φ =
∂y x r sin θ
∂φ
=0
∂z
h̄ ∂ 1 ∂ cos φ ∂
Lz = {r sin θ cos θ[sin θ sin φ + cos θ sin φ + ]
i ∂r r ∂θ r sin θ ∂φ
∂ 1 ∂ sin φ ∂
−r sin θ sin φ[sin θ cos φ + cos θ cos φ − ]
∂r r ∂θ r sin θ ∂φ
h̄ ∂
... Lz =
i ∂φ
h̄ ∂ 1 ∂
Lx = {r sin θ sin φ[cos θ − sin θ ]
i ∂r r ∂θ
∂ 1 ∂ cos φ ∂
−r cos θ[sin θ sin φ + cos θ sin φ + ]}
∂r r ∂θ r sin θ ∂φ
h̄ ∂ ∂
... Lx = {− sin φ − cot θ cos φ }
i ∂θ ∂φ
CHAPTER 4. MOTION IN THREE DIMENSIONS 101

h̄ ∂ 1 ∂ sin φ ∂
Ly = {r cos θ[sin θ cos φ + cos θ cos φ − ]
i ∂r r ∂θ r sin θ ∂φ
∂ 1 ∂
−r sin θ cos φ[cos θ − sin θ ]}
∂r r ∂θ
h̄ ∂ ∂
... Ly = {cos φ − cot θ sin φ }
i ∂θ ∂φ
h̄ ∂ ∂
L± = Lx ± iLy = {(− sin φ ± i cos φ) − cot θ(cos φ ± i sin φ) }
i ∂θ ∂φ
∂ ∂
= ±h̄e±iφ { ± i cot θ }
∂θ ∂φ
L2 = L2x + L2y + L2z = L+ L− + L2z − h̄Lz

∂ ∂ ∂ ∂ ∂2 h̄2 ∂
L2 = −h̄2 e+iφ [ + i cot θ ]e−iφ [ − i cot θ ] − h̄2 2 −
∂θ ∂φ ∂θ ∂φ ∂φ i ∂φ
L2 ∂ ∂ ∂ ∂ ∂2 1 ∂
2 = −[ ∂θ + i cot θ ∂φ + cot θ][ ∂θ − i cot θ ∂φ ] − ∂φ2 − i ∂φ

L2 ∂2 ∂ ∂ ∂ ∂ 2 ∂2 ∂
− 2 = 2
− i cot θ + i cot θ + cot θ 2
+ cot θ
h̄ ∂θ ∂θ ∂φ ∂φ ∂θ ∂φ ∂θ
∂ ∂2 1 ∂
−i cot2 θ + +
∂φ ∂φ2 i ∂φ
∂2 i ∂ ∂2 ∂ i cos2 θ ∂ ∂ ∂2
= 2
+ 2 + cot2 θ 2 + cot θ − 2 −i +
∂θ sin θ ∂φ ∂φ ∂θ sin θ ∂φ ∂φ ∂φ2
L2 ∂2 ∂ 1 ∂2 1 ∂ ∂ 1 ∂2
− = + cot θ + = sin θ +
h̄2 ∂θ2 ∂θ sin2 θ ∂φ2 sin θ ∂θ ∂θ sin2 θ ∂φ2
You see why I won’t discuss this in public!

4.5.1 Commutation Rules for Angular Momentum


The following commutators have been deduced starting from the definition of orbital angular mo-
mentum L ~ = ~r × p~ and the basic commutator [p, r] = h̄/i :

~ =L
[Lx , Ly ] = ih̄Lz or ih̄L ~ ×L
~

[L2 , Lz ] = 0
[L2 , L± ] = 0
Later these rules will be deduced from a more general point of view; namely, by studying the rotation
group. Two more types of commutator are also determined from the above cases:

[L± , Lz ] = [Lx , Lz ] ± i[Ly , Lz ] = −ih̄Ly ± iih̄Lx


= ∓h̄[Lx ± ih̄Ly ] = ∓L± h̄
and
[L+ , L− ] = [Lx + iLy , Lx − iLy ] = −ih̄iLz + i(−h̄i)Lz = 2h̄Lz ,
which simply replace the [Lx , Ly ] = ih̄Lz etc. forms.
We have therefore:
[L+ , L− ] = 2h̄Lz ; [L± , Lz ] = ∓h̄L± ;
[L2 , Lz ] = 0 and [L2 , L± ] = 0.
These have been organized to make application to the eigenvalue problems easier. The eigenvalue
problems deduced by separation of the wave function ψ = R(r)Y (θφ) are
CHAPTER 4. MOTION IN THREE DIMENSIONS 102

L2 Y`m (θφ) = λY`m (θφ) and Lz Y`m (θφ) = h̄mY`m (θφ)


= h̄2 `(` + 1)Y`m (θφ)

Y`m (θφ) = N P`m (cos θ)eimφ / 2π
The result that λ = h̄2 `(` + 1) has been anticipated by our notation, but it remains to show
that ` is indeed an integer and that m = 0, ±1, · · · ± `. That is we still have not shown that angular
momenta occur in discrete values only.

4.5.2 Dirac Notation for Spherical Harmonics


For convenience in completing the above proof and for later generalization to 21 integer spins, let us
now briefly consider the Dirac notation applied to the spherical Harmonics Y`m (θφ). The abstract
vectors in Hilbert space are defined with their projections giving the functions Y`m .
< θφ | `m >=< r̂ | `m >= Y`m (θφ)

< `m | θφ >=< `m | r̂ >= Y`m (θφ) .
Using this notation the orthonormal property of the Y`m0 s is:
Z
0 0
< ` m | ` m >= δ`` δmm =
0 0 < `m | r̂ > dr̂ < r̂ | `0 m0 >
Z

= Y`m (θφ) Y`0 m0 (θφ) dΩ r̂ .
R
This can be written simply as | r̂ >< r̂ | dΩ r̂ = 1. The action of an operator in an eigenvalue
equation can also be given formal expression as:
Lz | `m >= h̄ m | `m >
which means
Z
< r̂ | Lz | r̂0 > dΩr̂0 < r̂0 | `m >= h̄m < r̂ | `m >= h̄m Y`m (r̂)
Z
h̄ ∂
< r̂ | r̂0 > dΩr̂0 Y`m (r̂0 ) = h̄m Y`m (r̂).
i ∂φ
h̄ ∂
Y`m (θφ) = h̄m Y`m (θφ)
i ∂φ
The last steps involve using the explicit differential operator form of
h̄ ∂
< r̂ | Lz | r̂0 >= < r̂ | r̂0 >
i ∂φ
with similar equations holding for L2 , L± , etc.
The symbol < r̂ | r̂0 > is simple a Dirac delta function of the type
δ(θ − θ0 )
< r̂ | r̂0 >= δ(Ω − Ω0 ) = δ(φ − φ0 ) = δ(φ − φ0 )δ(cos θ − cos θ0 )
sin θ
with Z Z Z
dθ dφ sin θ < r̂ | r̂0 >= d cos θ dφ < r̂ | r̂0 >= dΩr̂0 δ(Ω − Ω0 ) = 1.

The formal expressions of Legendre’s equation then is simply


L2 | `m >= h̄2 `(` + 1) | `m >,
which follows using the above rules with
−h̄2 ∂ ∂ ∂2
< r̂ | L2 | r̂0 >= [ sin θ sin θ + ] < r̂ | r̂0 > .
sin2 θ ∂θ ∂θ ∂φ2
We can now proceed to prove that ` = integer and | m |≤ `.
CHAPTER 4. MOTION IN THREE DIMENSIONS 103

4.6 Quantization of Angular Momentum.


That only discrete values of ` and m are allowed will now be shown to be a consequence of the
orbital angular momentum commutators. Since these commutators arise from the basic commutator
[p, r] = h̄/i, which is an expression of the uncertainty principle, we see that the quantization of
angular momentum is understood as a direct consequence of our new understanding of the limitations
of our ability to measure. That is clearly an impressive insight compared to the seemingly arbitrary
rules used in the old quantum theory and presented even today in some elementary text books
(vector model).
The eigenvalue equation for L2 is of the form

L2 | `m >= h̄2 `(` + 1) | `m >= λ | `m >,


where ` is so far any real number. (L2 is Hermitian.) We also have

Lz | `m >= h̄m | `m >,

where m is any integer m = 0, ±1, ±2 · · · , if we require single valued wave functions. The commutator
[L2 , L± ] = 0 tells us that

L2 (L± | `m >) = L± L2 | `m >= h̄2 `(` + 1)(L± | `m >)

i.e. L± | `m > has the same eigenvalue for L2 of h̄2 `(` + 1). The commutator [Lz , L± ] = ±h̄L± tells
us that
Lz (L± | `m >) = L± Lz | `m > ±h̄L± | `m >= (m ± 1)h̄(L± | `m >),
i.e. L± | `m > has the eigenvalues (m ± 1)h̄ for the operator Lz . Thus we see the operators L± do
act as raising and lower operators on the quantum number m.
We can keep raising or lowering the eigenvalues of Lz using L± repeatedly; for example, after n
steps we have
(L± )n | ` m >= . | ` m ± n >
The question arises: is there some limit to this stepping process? That limit is provided by the fact
that L2 satisfies
L+ L− + L− L+
L2 = L2x + L2y + L2z = + L2z = L+ L− + L2z − h̄Lz
2
= L− L+ + L2z + h̄Lz
Recall [L+ , L− ] = 2h̄Lz . The quantity
Z Z
∗ ∗
(L− Y`m ) L− Y`m dΩ = Y`m L+ L− Y`m dΩ

=< `m | L+ L− | `m > ≥ 0
and also
< `m | L− L+ | `m > ≥ 0
are positive.
From these equations, it follows that

< `m | L2 − L2z + h̄Lz | `m >= h̄2 (`(` + 1) − m(m − 1)) ≥ 0

< `m | L2 − L2z − h̄Lz | `m >= h̄2 (`(` + 1) − m(m + 1)) ≥ 0


or
m(m ± 1) ≤ `(` + 1)
which means m is bounded.
Another result is
L− L+ + L+ L−
< `m | | `m >= h̄2 (`(` + 1) − m2 ) ≥ 0
2
CHAPTER 4. MOTION IN THREE DIMENSIONS 104

or
m2 ≤ `(` + 1).
(The equal sign is never realized)
Since m is bounded, for some maximum value

m ≡ mmax ≡ m> ,

no further increase is possible. For the state arrived at after n steps:

(L+ )n | `m >= . | `m> >

m> = m + n
for integer n steps up and similarly
m< = m − n0
for integer n0 steps down. We have
L+ | `m> >= 0
and
Lz | `m> >= h̄m> | `m> >
and

L2 | `m> >= h̄2 `(` + 1) | `m> >= [L− L+ + h̄2 (m> (m > +1)] | `m> >= h̄2 m> (m> + 1) | `m> > .

We find therefore `(` + 1) = m> (m> + 1). Similar stepping down leads to a lowest m-value m< =
m − n0 and `(` + 1) = m< (m< − 1). Combining these, we get

m> (m> + 1) = m< (m< − 1)

(m> + m< )(m> − m> + 1) = m> (m> + 1) − m< (m< − 1) − m> m< + m< m> = 0.
Hence
m> − m< + 1 > 0
and we must have:
m> = −m<
... m> − m< = 2m> = n + m − m − n0 = n − n0 = an integer
are since integer steps are used in the raising and lowering process. Using `(` + 1) = m> (m> + 1),
it follows that ` = m> and from `(` + 1) = m< (m< − 1) we get ` = m> . The cases ` = −m> − 1
and ` = m< − 1 are rejected since ` ≥ 0. We arrive at the conclusion:

2m> = integer = 2`
1
... ` = integer or integer.
2
Also
m ≤ m> ... m ≤ `; m = 0, ±1, ±2, · · · ± `.
We will rule out the 21 integer values for orbital angular momentum in a moment based on orbital
angular momentum having a differential operator realization.
CHAPTER 4. MOTION IN THREE DIMENSIONS 105

4.6.1 Explicit Representation of L+ Step:


We can express the property L± | ` ± ` >= 0, for orbital angular momentum using the explicit
differential operator representation in r-space as
∂ ∂ ∂
( ± i cot θ )Y`±` (θϕ) = 0 = ( − ` cot θ)e±i`ϕ P`±` = 0
∂θ ∂ϕ ∂θ

using Y`` ∼ e±i`φ P`±` . The solution to the above relation is of the form

P`±` = c±
` sin` θ

and
Y`±` = c± `
` sin θ e
±i`ϕ
.
All properties of the P`m can be obtained using explicit representation of the stepping process and
this solution. Also we can rule out 21 integer `-values since they would be inconsistent with

L− Y 12 12 = Y 12 − 21

since with √
Y 12 21 = c± sin θe±iϕ/2
the operator
∂ ∂
L− → ± i cot θ
∂θ ∂ϕ
won’t yield that result. e Hence, for operators defined in the r-representation, i.e. orbital angular
momentum 12 is not allowed. For spin we will later deal with a new space - spin - space, which is
not defined in an r-representation.

4.7 The Radial Equation.


Having discussed the angular part of our separated Schrödinger equation, let us now examine the
radial equation
p2 h̄2 `(` + 1)
[ r + V (r) + ]Rn` (r) = En` Rn` (r)
2µ 2µ r2
Note that the radial wave function and the eigenvalue are independent of m for a central potential.
That fact yields a degeneracy 2` + 1 (= number of allowed m-values for a given `) ; i.e. 2` + 1
different solutions for the same energy value. The 3-D wave function

ψ = Rn` (r) Y`m (θφ) = (un` /r) Y`m (θφ)


2
has been separated to obtain three one-dimensional equations. Using the form p2r = −h̄2 1r ∂r

2 r , we

have
h̄2 d2 h̄2 `(` + 1)
[− 2
+ + V (r)]un` = En` un` (r)
2µ dr 2µ r2
or
d2 `(` + 1) 2µ 2µ
{− 2 + [ + 2 V (r)]}un` = 2 En` un` (r)
dr r2 h̄ h̄
2
= kn` un` (r)
h̄2 2
En` ≡ k .
2µ n`
The boundary conditions on un` and Rn` follow from our requirement that the norm be bounded
for bound states where Z ∞ Z ∞
dr u∗n` un` = 1 = dr r2 | Rn` |2 .
0 0
e In addition,Y 1 − 1 ∝ sin−1/2 θ is singular at zero angle.
2 2
CHAPTER 4. MOTION IN THREE DIMENSIONS 106

Thus we must have Rn` and un` → 0 at r = ∞. For a bound state

h̄2 2
kn` = ikB or En` = − k <0
2µ B
and the solution is for large r
un` → e−kB r .
The Hermiticity condition (see earlier home work) implies that
Z ∞ Z ∞
2 ∗
drr Rn` pr Rn` = dr r2 (p†r Rn` )∗ Rn`
0 0

with p†r = pr only if the surface terms rRn` vanish at the boundaries.f Therefore, we also require that
Rn` be finite at r = 0 and hence that un` (0) = 0! The additional boundary conditions that R, R0 and
u, u0 be finite, continuous, and single valued again arise from our conditions on the physical currents
and densities.
The above radial equation along with the boundary conditions is a Sturm-Liouville problem
for which one can prove that a complete orthonormal set is generated. We thus have a complete
orthonormal set of radial wave functions generated for each central Ppotential V(r). Of course, one
must include both the continuum and bound solutions. We have α cα uα (r) = any function of r
(i.e. which is finite and single-valued etc.) and we can express these properties using orthonormality
(ON) Z ∞
u∗α (r)uα0 (r) dr = δαα0
0
g
and closure (C)
X 1
uα (r) u∗α (r0 ) = δ(r − r0 ),
α
r2
including bound and continuum states.

4.7.1 Spherical Bessel Functions - Free Radial Waves.


One complete orthonormal set of radial wave functions which are of particular importance are the
solutions with V = 0; these are the spherical Bessel functions. The free particle solutions are thus
generated for each `, m value. Since for V = 0 no bound states occur, we have

h̄2 2
E= k ≥0
2µ 0
only and with k0 r ≡ ρ, the radial equation is with

p2r h̄2 1 ∂ 2 ∂ h̄2 k02 ∂ 2 ∂ 1 ∂ 2 ∂


=− 2
r =⇒ − 2
ρ = −E( 2 ρ )
2µ 2µ r ∂r ∂r 2µ ρ ∂ρ ∂ρ ρ ∂ρ ∂ρ
and therefore
1 ∂ 2 ∂ `(` + 1)
[+ ρ − + 1]R` (ρ) = 0.
ρ2 ∂ρ ∂ρ ρ2
The solution of this equation, which satisfies the condition that R` (0) be finite and u` (0) = 0 are
the spherical Bessel functions, which are given in terms of Bessel functions by:
r
π
j` (ρ) = J 1 (ρ).
2ρ `+ 2

These functions have the important properties:


1 π
j` (ρ) → sin(ρ − ` ) + O(ρ−2 ) for ρ → ∞
ρ 2
f For scattering problems we can use box normalization to assure hermicity.
g Recall that completeness plus orthonormality lead to the closure form
CHAPTER 4. MOTION IN THREE DIMENSIONS 107

ρ`
j` (ρ) → for ρ → 0.
(2` + 1)!!
Note the double factorial is: (2` + 1)!! = (2` + 1)(2` − 1) · · · 3 · 1.
The irregular solutions, which are useful away from the r = 0 point where they are singular, are
called the Neumann functions, whose properies include:
r
π
n` (ρ) = N 1 (ρ). ,
2ρ `+ 2
1 π
n` (ρ) → − cos(ρ − ` ) + O(ρ−2 ) for ρ → ∞
ρ 2
1
n` (ρ) → (2` − 1)!! ( )`+1 for ρ → 0.
ρ
Additional combinations that will be useful later are the Hankel functions (first and second kind)

eiρ
h` (ρ) = j` (ρ) + in` (ρ) → for ρ → ∞,
i`+1 ρ
and h∗` (ρ) = j` (ρ) − in` (ρ),
The property of being a complete orthonormal set of continuum radial wave functions are ex-
pressed by:
2 ∞
Z
1
j` (k 0 r) j` (kr) r2 dr = 2 δ(k − k 0 )
π 0 k
and Z ∞
2 1
j` (kr0 ) j` (kr) k 2 dk = δ(r − r0 ),
π 0 r2
h
which are the orthonormality and closure properties.

4.8 Spherical Harmonics - Complete Set of Angle Functions.


The angular equations also generate a complete orthonormal set. These are the spherical harmonics
Y`m (θφ) which are a complete, orthonormal (CON) set of functions which satisfy the requirements
of being finite, single-valued and continuous. Any such angle dependent function can be expanded
in terms of these functions using a combination
∞ X
X `
F(θφ) = c`m Y`m (θφ).
`=0 m=−`

Since Z
Y`∗0 m0 (θφ) Y`m (θφ)dΩ = δ``0 δmm0 (ON ),

we have Z

c`m = Y`m (θφ) dΩ F(θφ),

and again the completeness can be expressed in the closure form


X
∗ δ(θ − θ0 )
Y`m (θφ) Y`m (θ0 φ0 ) = δ(Ω − Ω0 ) = δ(φ − φ0 )
sin θ
`m

= δ(φ − φ0 )δ(cos θ − cos θ0 )


= δ(r̂ − r̂ 0 ) .
h In some cases, the basis is extended to include angular dependence and an energy normalization:
2 dk 1/2
< r̂ | E`m >= ( k2 ) j` (kr) Y`m (r̂)
π dE
CHAPTER 4. MOTION IN THREE DIMENSIONS 108

The formal version of these relations are


X
| `m >< `m |= I
`m

and
< `0 m0 | `m >= δ``0 δmm0 .
Note in the above cases a sum over both ` and m occurs. Also of interest is the case of the addition
theorem in which only a sum on m occurs
X
∗ 2` + 1
Y`m (θ1 φ1 )Y`m (θ2 φ2 ) = P` (cos θ12 ),
m

(see later for a proof using rotation group properties).


Now any general wave function can be expanded using the complete set of radial and angular
wave functions and a combination of the stationary states
i
X
Ψ(~r, t) = cn`m e− h̄ En` t Rn` (r) Y`m (θφ),
n`m

where n labels both discrete and continuum eigenstates.Here Ψ(~r, t), a general solution of HΨ =
ih̄ ∂Ψ/∂t, is expanded using the stationary state solutions of H0 + V.
~
As a particularly useful example let us consider the free wave solution eik·~r /(2π)3/2 expanded in
the complete orthonormal set j` (kr)Y`m (θφ).
~
X
eik·~r = c`m Y`m (θr φr )j` (kr) = eikr cos θkr
`m


using the j` radial solutions for V = 0. One can prove that c`m = Y`m (θk φk )4πi` [either by comparing
series expansions or using the integral representation of j` (kr)]. Hence the plane wave can be
expanded as
~ r
eik·~r 2 X ` ∗
3/2
= i j` (kr) Y`m (k̂) Y`m (r̂)
(2π) π
`m

or using the addition theorem


~
eik·~r 1 X
= i` j` (kr)(2` + 1)P` (k̂ · r̂) .
(2π)3/2 (2π)3/2 `

We will use this later for scattering theory; for now let me illustrate the use of these relations in a
way that will be important (see the scattering amplitude later in the Born approximation.)

Example.
From
~ r
eik·~r 2X `
< ~r | ~k >= = ∗
i j` (kr) Y`m (r̂) Y`m (k̂),
(2π)3/2 π
`,m

we deduce
~ ~0 0
e−ik·~r eik ·~r
Z
< ~k | V | ~k 0 >= < ~
r | V | ~
r 0
> d~r d~r 0
Nonlocal
(2π)3/2 (2π)3/2
Z
1 ~ ~0
= 3
e−ik·~r V (| ~r |)eik ·~r d~r Local, central
(2π)
For a local, central potential:

(~r | V | ~r 0 ) ≡ V (| ~r |)δ(~r − ~r 0 )
Z ∞
2 X X −`
< ~k | V | ~k 0 >= i Y`m (k̂) j` (kr) Y`m∗
(r̂) V (r) r2 dr
π 0 0 0
`m ` m
CHAPTER 4. MOTION IN THREE DIMENSIONS 109

0
j`0 (k 0 r) Y`0 m0 (r̂) i` dΩr̂ Y`∗0 m0 (k̂ 0 ),
Using Z

Y`m (r̂) Y`0 m0 (r̂) dΩ( r̂) = δ``0 δmm0 ,

the above reduces to X


< ~k | V | ~k 0 >= (k | v` | k 0 )Y`m (k̂)Y`m

(k̂ 0 )
`m

1 X
= (k | v` | k 0 ) (2` + 1) P` (k̂ 0 · k̂),

`

with k̂ 0 · k̂ = cos θ and Z ∞


0 2
(k | v` | k ) = j` (kr) r2 dr V (r) j` (k 0 r).
π 0
See later for application of this to partial wave expansions in scattering.
Chapter 5

Bound States

5.1 The Hydrogen Atom.


We can now consider solutions of the radial equation with V 6= 0. An important case is a hydrogen-
like atom (hydrogenic), which provides a very useful complete orthonormal set of wave functions.
The problem of an electron moving in a Coulomb field −Ze2 /r is of great historic importance. It
provided a means of understanding Bohr’s model and going beyond it to understanding intensities,
selection rules, scattering, etc. Schrödinger’s equation provided great insight into the dynamics, but
in fact the Heisenberg or matrix methods were developed first, so we are turning history around. I
suggest you look at M. Jammers’ book for a discussion of the historic development. Pauli’s Handbuch
der Physik article has I believe a matrix mechanics version of the hydrogen atom derivation
For our present discussion, the wave functions of hydrogenic atoms provide not only a solution
to the hydrogen problem but a complete set that can be (and is) used to treat atoms with many
electrons. (These studies involve the Hartree-Fock or Thomas Fermi methods and more advanced
many-body techniques). Other applications of hydrogenic wave functions occur in the processes of
mesonic and antiprotonic π − , µ− , K − and p̄ atoms and capture, and also for the short range part of
the quark-antiquark interaction.
Let us examine the bound state for V = −Ze2 /r now and consider the continuum states later.

5.1.1 The Bound State - Radial Equation.


Instead of using more elegant methods (such as raising and lowering operators) to solve the radial
equations, let me first present the standard series solution method and its relation to associated
Laguerre polynomials. The bound state solution for bound state boundary conditions requires we
find bounded solutions of
h̄2 d2 h̄2 `(` + 1) Ze2
[− + − − En` ]un` (r) = 0,
2µ dr2 2µ r2 r

where the reduced mass µ = me Mp /(me + Mp ) ≈ me . The units for this problem are those given by
the Bohr theory. We assume µ ' m ' electron mass; (it can be replaced by the reduced mass)

Z 2 e2 1 1 1 1 13.6ev
En` = − 2
= − Z 2 α2 mc2 2 = − Z 2
2a0 n 2 n 2 n2
2

with a0 = me 2 = .53Å. (See H.W. #1). So far, n is any real number, it will be shown later to be an

integer. The distance is measured by


a0
r= nρ.
2Z
changing variables E → n and r → ρ, we have

4Z 2 d2 `(` + 1) 2 2m 2Z 2 e2 Z 2 e2 1 2m
un` (ρ) + [− 4Z + − ]un` (ρ) = 0
a20 n2 dρ2 a20 n2 ρ2 h̄2 a0 nρ 2a0 n2 h̄2
CHAPTER 5. BOUND STATES 111

or
d2 un` `(` + 1) n 1
2
+ [− + − ]un` (ρ) = 0.
dρ ρ2 ρ 4
Now for ρ → ∞ this becomes u00 = 41 u with an asymptotic solution e±ρ/2 . Since u → 0 is required
for a bound state we must have only e−ρ/2 for large ρ. (Note: ρ depends implicitly on n and care
must be exercised in proving orthogonality for different n-values.)
Based on the asymptotic behavior, we take

u = e−ρ/2 F (ρ)

and deduce
1
u0 = − u + e−ρ/2 F 0
2
1 1 1
u00 = − u0 − e−ρ/2 F 0 + e−ρ/2 F 00 = + u − e−ρ/2 F 0 + e−ρ/2 F 00
2 2 4
n `(` + 1)
−F 0 + F 00 + [ − ]F = 0.
ρ ρ2
Now take F as a power series in ρ of the form
X∞
F =( as ρs )ρλ
s=0

where we have separated out a term ρλ with λ > 0 in order to keep u(0) = 0.

X
F0 = as (s + λ)ρs+λ−1
s=0


X
F 00 = as (s + λ)(s + λ − 1)ρs+λ−2
s=0

combining

X
as (s + λ)(s + λ − 1)ρs+λ−2 − as (s + λ)ρs+λ−1 + nas ρs+λ−1
s=0

−`(` + 1)as ρs+λ−2 ] = 0



X
ρλ−2 a0 [λ(λ − 1) − `(` + 1)] + nas ρs+λ−1 − as (s + λ)ρs+λ−1
0

X
+ [as (s + λ)(s + λ − 1)ρs+λ−2 − `(` + 1)as ρs+λ−2 ] = 0
s=1

X
= ρλ−2 a0 [λ(λ − 1) − `(` + 1)] + ρs+λ−1 [(n − s − λ)as + as+1 (s + 1 + λ)(s + λ) − `(` + 1)as+1 ] = 0
s=0

Each power of ρ has its coefficient ≡ 0!

... λ(λ − 1) = `(` + 1)

(n − s − λ)as = +[`(` + 1) − (s + 1 + λ)(s + λ)]as+1


First equation is solved by λ = ` + 1 or λ = −`; λ = −` is rejected since it gives F (0) ⇒ ∞. If we
took a0 = 0 λ = −`, we would get F (ρ) = 0 everywhere using second equation ... need a0 6= 0 and
λ = ` + 1. The recurrence relation with a0 6= 0 is:

[n − s − ` − 1] [n − s − ` − 1]
as+1 = as = − as
`(` + 1) − (s + ` + 2)(s + ` + 1) (2 + s)(s + ` + 1) + `s
CHAPTER 5. BOUND STATES 112

For large s (s >> `, s >> n),


as+1 s 1
→ 2 →
a s s
and the series is

X X ρs
as ρs → .
0
s!
1/(s+1)! 1
P ρs
Since 1/s! ⇒ s and s! = eρ , we get for large ρ, where the large s-values determine the sum

F (ρ) → eρ ρλ
and
lim u(ρ) → eρ/2 ρλ → ∞ .
ρ→∞

Since we require u(ρ) be bounded in its norm this divergence must be prevented by truncating the
series at some value of s. The value of n is therefore restricted, and hence only certain eigenvalues
are allowed by the boundary conditions on u(ρ). This truncation of a series solution to satisfy a
boundary condition is the typical manner in which boundary conditions lead to discrete quantum
numbers.
Setting n = integer ≥ ` + 1, then n = s + ` + P1 for some value of s and as+1 = 0 by virtue
of the recurrence relation given above. The sum 0 as ρs is therefore a polynomial with order of
(n − ` − 1) = smax = order of the polynomial. [Note that n > `(` + 1) 6= n(n + 1) which prevents a
zero in the denominator of the recurrence relation. ] Our result is
n−`−1
X
un` (ρ) = e−ρ/2 ρ`+1 0 −ρ/2 `+1
as ρs = Nn` e ρ L(ρ)
s=0

where n = integer ≥ ` + 1 and


n−s−`−1
as+1 = as .
`(` + 1) − (` + 2 + s)(` + 1 + s)

Atomic Convention.
The atomic convention for the principal quantum number n is that:
n − ` − 1 = number of nodes in un` (ρ) not counting the ρ = 0 and ρ = ∞ nodes.
The factors ρ`+1 and e−ρ/2 account for the zeroes at ρ = 0 and ρ = ∞ respectively; whereas, the
Pn−`−1
polynomial L(ρ) ≡ 0 as ρs is of order n − ` − 1 with real coefficients and hence has n − ` − 1
zeroes (or nodes).

Laguerre polynomials.
Pn−`−1
The sum s=0 as ρs is within a normalization N 0 , simply the associated Laguerre polynomials
(see NBS (A&S) book for series expansion of Laguerre polynomials). To check this fact note that
u = eρ/2 ρ`+1 L(ρ) used in
n `(` + 1) 1
u00 + [ − − ]u = 0
ρ ρ2 4
yields
n `(` + 1)
F 00 − F 0 + [ − ]u − 0
ρ ρ2
with
F = ρ`+1 L(ρ)

ρL00 + [(2` + 1) + 1 − ρ]L0 + (n + ` − (2` + 1))L = 0,


which is Laguerre’s equation whose solution is labelled as Laσ with
d2 a d
ρ L (ρ) + (a + 1 − ρ) Laσ (ρ) + (σ − a)Laσ (ρ) = 0,
dρ2 σ dρ
and a = 2` + 1 and σ = n + ` for Hydrogenic. Note:
σ − a = n − ` − 1 = # nodes
CHAPTER 5. BOUND STATES 113

5.2 Wave Function For Hydrogen.


We finally obtain the solution for V = −Ze2 /r as:

ψn`m (~r ) = −Nn` e−ρ/2 ρ` L2`+1


n+` (ρ) Y`m (θφ) (5.1)
Recall that u = rR; ψ = RY`m
r
ρ= a0
2Z n

1 mc2
En` = − Z 2 α2 2
2 n
n ≥ ` + 1 = integer
where n − ` − 1 = number of nodes in ψ except for those at r = 0, ∞.

Level Scheme for Hydrogenic Atoms.


Using spectroscopic notation ` = 0, 1, 2, 3 · · · → s, p, d, f · · · ,

` = 0, n = 1, 2, 3 · · · → 1s, 2s, 3s · · ·
` = 1, n = 2, 3, 4 · · · → 2p, 3p, 4p · · ·
` = 2, n = 3, 4, 5 · · · → 3d, 4d, 5d · · ·
Here 1s, 2p, 3d have zero nodes, 2s, 3p, 4d have one node and 3s, 4p, 5d have two nodes in their
wavefunction.

p
n=5 5s p 5p 5dp 5fp Degeneracy=25
p p p
3 Nodes p p p
p p p p p
n=4 4sp p 4pp p 4dp p 4f Degeneracy=16
p p p
p p
p p p
p p p
2 Nodes p p
p p p
p p
p p
n=3 3sp p 3pp p 3d p Degeneracy=9
p p
p p
p p
p
1 Nodes
p
p p
p p
p p
0 Nodes
p p
p p
p p
n=2 2sp 2pp Degeneracy=4
p
p
p
p
p
p
p
p
1sp p
n=1 Ground State Degeneracy=1

Figure 5.1: Level Scheme for Hydrogenic Atoms. The dotted


Pn−1curves join states with the same number
of nodes, as indicated in small print. The degeneracies `=0 (2` + 1) = n2 are also indicated on the
right side.
CHAPTER 5. BOUND STATES 114

5.2.1 Normalization - Generating Functions


The normalization factor still requires discussion. For that determination and also for finding matrix
elements for other corrections and effects (such as the relativistic and spin-orbit corrections and
dipole transitions), we need to develop means of evaluating matrix elements using ψ. The generating
function technique is a useful way of handling such problems. The generating function for the
Laguerre polynomials, which serves to define the polynomials used here, is

e−ρt/1−t X
a tσ−a
Ga (ρ, t) = = L σ (ρ) (−1)a
(1 − t)a+1 Γ(σ + 1)
σ≥a

This is equivalent to our earlier definition using Laguerre’s equation since


∂ ∂ ∂2
[t + (a + 1 − ρ) +ρ ] Ga (ρ, t) = 0
∂t ∂ρ ∂ρ2
X (−1)a 00 0
= {ρLσa + (a + 1 − ρ)Lσa + (σ − a) Laσ }tσ−a
σ>a
Γ(σ + 1)
00 0
or ρ Lσa + (a + 1 − ρ)Lσa + (σ − a)Laσ = 0 as before; with σ = n + ` and a = 2` + 1.
Our reason for introducing this generating function is to enable us to find the normalization Nn`
for any state. In the appendix, we use Ga (ρ, t) to show that
2Z 3/2 (n − ` − 1)! 1 −ρ/2 ` 2`+1
ψn`m (~r) = −( ) [ ]2 e ρ Ln+` (ρ) Y`m (r̂).
a0 n 2n (n + `)!3
For example, this reduces to a simpler form for n = ` + 1 (no nodes wave function) i.e.
2Z 3/2 1 1
ψn`m (~r ) = −( ) [ ] 2 ρ` e−ρ/2 (−1)a (2n − 1)! Y`m
a0 n 2n(2n − 1)!3
2Z 3/2 1 1
= +1( ) [ ] 2 ρ` e−ρ/2 Y`m (θφ),
a0 n 2n(2n − 1)!
(−1)a t0 La
using the fact that G2`+1 (ρ, 0) = 1 = σ!
σ
for σ − a = n − ` − 1 = 0,
... L2`+1
n+1 (ρ) = (−1)
2`+1
σ!
= −(2n − 1)!
Nodeless wave functions are
2Z 3 1 1
ψn`m (~r ) = ( )2 ( ) 2 ρ` e−ρ/2 Y`m (θφ).
a0 n 2n(2n − 1)!

Appendix - Determination of Normalization


Normalization Nn` is determined for ψn`m = −Nn` e−ρ/2 ρ` L2`+1 n+` (ρ) Y`m (r̂) by taking
Z Z ∞
d~r | ψn`m |2 = 1 = Nn`
2
dr r2 e−ρ ρ2` | Laσ (ρ ) |2 =
0
Z ∞
a0
2
Nn` ( n)3 dρ ρ2`+2 e−ρ | Laσ (ρ) |2
2Z 0
σ = n + `; a = 2` + 1; σ − a = n − ` − 1 > 0.
Need ∞
σ!3
Z
dρ ρa+1 e−ρ | Laσ (ρ) |2 = (2σ − a + 1)
0 (σ − a)!
2n(n + `)!3
=
(n − ` − 1)!
can prove this using generating function. Thus
2 2Z 3 (n − ` − 1)!
Nn` =( ) .
a0 n 2n[(n + `)!]3
CHAPTER 5. BOUND STATES 115

5.3 Harmonic Oscillator in One, Two and Three Dimensions


5.3.1 The Radial Differential Equation
For the three dimensional, isotropic a harmonic oscillator we have the radial equation in spherical
coordinates
h̄2 d2 h̄2 `(` + 1) 1
[− + + mω 2 r2 ]un` = En` un` (r) (5.2)
2m dr2 2m r2 2
or
d2 `(` + 1) 2µ 1 2µ
{− 2 + [ 2
+ 2 mω 2 r2 ]}un` = 2 En` un` (r)
dr r h̄ 2 h̄
2
= kn` un` (r). (5.3)
un` (r)
With ψ = Rn` (r)Y`m (θ, φ) = we have the bound state boundary condition un` → 0,
r Y`m (θ, φ),
q

for r = 0 and r = ∞. We introduce the length scale called the “oscillator length” b = mω and
ρ ≡ r/b, as the unitless distance. The above radial equation then converts to

h̄2 1 d2 h̄2 1 `(` + 1) 1


[− + + mω 2 b2 ρ2 ]un` = En` un` (r) (5.4)
2m b2 dρ2 2m b2 ρ2 2
and
d2 `(` + 1) 2m h̄
[− + + ρ2 ]un` = 2 En` un` = n` un` . (5.5)
dρ2 ρ2 h̄ mω
we have defined E = h̄ω2 n` .
d2 `(`+1) 2
For small values of ρ, the centrifugal term dominates and [− dρ 2 + ρ2 +ρ ]un` = n` un` reduces
2 `(`+1)
d s
to [− dρ 2 + ρ2 ]un` = 0. That low ρ region thus suggests a solution having the form ρ = ρ , since
u0 = sρs−1 , u00 = s(s − 1)ρs−2 = `(` + 1)ρs−2 . Note then that s(s − 1) = `(` + 1), which yields ` = −s
or s = ` + 1. To get u(0) = 0, we need s > 0 and since ` > 0 also, the only acceptable solution is
s = ` + 1. The ansatz for the solution, following the same type of procedures used for the hydrogenic
bound state is therefore
u(ρ) = ρ`+1 F (ρ),
The following steps are needed:

u0 (ρ) = (` + 1)ρ` F (ρ) + ρ`+1 F 0 (ρ)


00
u (ρ) = `(` + 1)ρ `−1
F (ρ) + 2(` + 1)ρ` F 0 (ρ) + ρ`+1 F 00 (ρ)
−[`(` + 1)ρ`−1 F (ρ) + 2(` + 1)ρ` F 0 (ρ) + ρ`+1 F 00 (ρ)] + `(` + 1)ρ`−1 F (ρ) + ρ`+3 F (ρ) = n` ρ`+1 F (ρ).
(5.6)

Clearing common factors, we get


`+1 0
−F 00 − 2 F + ρ2 F = n` F, (5.7)
ρ
1 2
which for large ρ is F 00 = ρ2 F, which gives the form F = e− 2 ρ ; note that F 0 = −ρF, F 00 =
−ρF 0 − F ≈ ρ2 F.
1 2
Our next ansatz is then F (ρ) = e− 2 ρ L(ρ), with L(ρ) to be determined by:
1 2 1 2
F0 = e− 2 ρ L0 − ρe− 2 ρ L
1 2 1 2 1 2 1 2
F 00 = −2ρe− 2 ρ L0 (ρ) + e− 2 ρ L00 (ρ) − e− 2 ρ L(ρ) + ρ2 e− 2 ρ L(ρ)
`+1 0
= −(n` − ρ2 )F − 2 F
ρ
1 2 ` + 1 − 1 ρ2 0 1 2
= (ρ2 − n` )e− 2 ρ L(ρ) − 2 (e 2 L − ρe− 2 ρ L). (5.8)
ρ
a By isotropic we mean that the spring constant is the same in all three directions, so that ωx = ωy = ωz = ω.
CHAPTER 5. BOUND STATES 116

Reordering terms gets us to


`+1
L00 + 2ρ( − 1)L0 + (n` − 2` − 3)L = 0,
ρ2
which can be cast into a Laguerre form

d2 a d
z L (z) + [a + 1 − z] Laσ (z) + (σ − a)Laσ (z) = 0,
dz 2 σ dz
by invoking the substitutions z = ρ2 , dz = 2ρdρ, · · · , to arrive at

d2 a 1 d 1
z L (z) + [ + ` + 1 − z] Laσ (z) + (n` − 2` − 3)Laσ (z) = 0,
dz 2 σ 2 dz 4
from which we identify
1
a=`+
2
1 1
(n` − 2` − 3) = σ − a = σ − ` − ,
4 2
with σ − a equal to the number of nodes in Laσ , which in the nuclear convention ≡ n − 1. We have
σ = a + n − 1 = n + ` − 1/2.
Once again we could generate a series solution and see that it needs to be truncated; whereby
a boundary condition requires discrete quantum numbers. Instead of belaboring that aspect, we
go directly to use of Laguerre polynomial solution which incorporate those considerations. The full
1 2
radial wavefunction is therefore u(ρ) = Nn` ρ`+1 e− 2 ρ Laσ (ρ), which has σ − a = n − 1 nodes, not
counting the nodes at ρ = 0 or ρ = ∞.
We can solve for the energy and obtain n` = 4n + 2` − 1 and
1
En` = h̄ω(2n + ` − ).
2
For the convention based on the number n0 = n−1 (now with n0 =number of nodes in ψ not counting
the 0, ∞ points), we get:
3
En` = h̄ω(2n0 + ` + ).
2
Note the zero point energy of h̄ω3/2, for the three degrees of freedom. So we have n = 1, 2, 3 . . .
and n0 = 0, 1, 2, · · · as the two conventions used to label the energy eigenstates of a 3-D harmonic
oscillator.
Our final resultant wave function is
r` 1 2
/b2 `+1/2
ψn`m = Nn` e− 2 ρ Ln+`−1/2 (r2 /b2 )Y`m (θ, φ). (5.9)
b`+1
Final Form of the 3-D Oscillator Wavefunction
The result is
1
(−1)`+ 2 2(n − 1)! r 1 r 2 `+1/2 r2
3/2
[ 3
]1/2 ( )` e− 2 ( b ) Ln+`−1/2 ( 2 ) Y`m (θ, φ). (5.10)
b Γ(n + ` = 1/2) b b
CHAPTER 5. BOUND STATES 117

n n0 Λ ` (n`) E/(h̄ω) = (Λ + 3/2)


1 0 0 0 1s 3/2
1 0 1 1 1p 5/2
1 0 2 2 1d 7/2
2 1 2 0 2s 7/2
1 0 3 3 1f 9/2
2 1 3 1 2p 9/2
1 0 4 4 1g 11/2
2 1 4 2 2d 11/2
3 2 4 0 3s 11/2
1 0 5 5 1h 13/2
2 1 5 3 2f 13/2
3 2 5 1 3p 13/2
1 0 6 6 1i 15/2
2 1 6 4 2g 15/2
3 2 6 2 3d 15/2
4 3 6 0 3s 15/2

Table 5.1: We use spectroscopic notation for ` = 0, 1, 2, 3, . . . , → s, p, d, f, g, h, i · · · . Also the label


Λ ≡ 2n0 + ` = 2n + ` − (1/2) is used.

5.3.2 The Differential Equation and Ladder Operators for 1-D Oscillator
Now consider a one dimensional harmonic oscillator. We shall first take a differential equation
viewpoint, then introduce ladder (raising & lowering) operators and then get to the elegant and
useful creation and annihilation operator formalism. At the end of this section we will consider
three oscillators in the three Cartesian direction and return to representations of the 3-D harmonic
oscillator.
The stationary-state, 1-D oscillator Schrödinger equation is:

h̄2 d2 1
[− + mω 2 x2 ]ψ(x) = En ψ(x).
2m dx2 2
p
Again defining an “oscillator length” by b = h̄/mω and x = bρ, the above equation is recast as:

d2
[− + ρ2 ]ψ(ρ) = n ψ(ρ)
dρ2
h̄ω
with En ≡ 2 n . This has a solution of the form
2
ψ = e−ρ /2
H(ρ),

which after it is inserted into the above equation gives

d2 d
H − 2ρ H + (n − 1)H = 0.
dρ2 dρ

This is Hermite’s equation b , which yields bounded Hermite polynomials Hn (ρ) if n − 1 = 2n, with
n = 0, 1, 2, · · · . Hence, the eigenvalue is
h̄ω 1
En = n = h̄ω(n + ).
2 2
The generating function for the Hermite polynomials is

2 X 1
G(z, t) = e2zt−t = Hn (z)tn .
n=0
n!
b See A & S page 781
CHAPTER 5. BOUND STATES 118

The Rodrigues’ formula for these polynomial is:


2 d n −z2
Hn (z) = (−1)n ez ( ) [e ].
dz
Using these, one can obtain the first four Hermite polynomials as H0 (z) = 1; H1 (z) = 2z; H2 (z) =
4z 2 − 2; H3 (z) = 8z 3 − 12x; H4 (z) = 12 − 48z 2 + 16z 4 .
Our 1-D oscillator wave function is
1 2 x 1 1 2 x
ψn (x) = Nn e− 2 (x/b) Hn ( ) = [ n √ ]1/2 e− 2 (x/b) Hn ( ),
b 2 b n! π b

where the normalization can be found from the generating function to be Nn = [ 2n n! 1b √


π
]1/2 .
The first few 1-D oscillator wave functions are:
1 1 2
ψ0 (x) = √ ]1/2
[ e− 2 (x/b)
20 b 0! π
1 1 2
ψ1 (x) = [ 1 √ ]1/2 e− 2 (x/b) (2x/b)
2 b 1! π
1 1 2
ψ2 (x) = [ 2 √ ]1/2 e− 2 (x/b) (4(x/b)2 − 2)
2 b 2! π
1 1 2
ψ3 (x) = [ 3 √ ]1/2 e− 2 (x/b) (8(x/b)3 − 12x/b) (5.11)
2 b 3! π
The generating function procedure, which can be used for many other cases is:
Z ∞ Z ∞
0 02
−x2 0 2 2
e G(x, t) G(x, t ) dx = e2xt−t e2xt −t e−x dx
−∞ −∞
Z ∞ X ∞
∞ X
1 1 0 2
= Hn (x)tn 0 Hn0 (x)t n e−x dx
−∞ n=0 n0 =0 n! n!
Z ∞ Z ∞

0 0 2
2tt0 −x2 0
e2tt e−(x−t−t ) dx = e e dx = e2tt π, (5.12)
−∞ −∞

hence
∞ ∞ X∞ Z ∞
X 1 √ X 1 1 n 0n 2
(tt0 )n [ 2n π ] = 0
t t [ Hn (x)Hn0 (x)e−x dx ] (5.13)
n=0
n! n=0 0
n! n ! −∞ n =0

This holds for all values of t and t0 and we arrive at:


Z ∞
2 √
Hn (x)Hn0 (x)e−x dx = δnn0 2n n! π
−∞

and Z ∞
2
Hn (x/b)Hn0 (x/b)e−(x/b) dx = δnn0 2n n! b .
−∞

Raising and Lowering Differential Operators


Note that we can introduce a raising differential operator by
1 1 ∂
a† = √ [x − b2 ]
2b ∂x
1 1 i
= √ [x − px ]. (5.14)
2b mω
q
Recall that b = h̄
mω . The lowering operator is the adjoint of a† or

1 1 ∂
a = √ [x + b2 ]
2 b ∂x
1 1 i
= √ [x + px ]. (5.15)
2b mω
CHAPTER 5. BOUND STATES 119

These operators connect one oscillator state to the next higher or lower, which follows from
1 1 ∂ 1 1 2
a† ψn = √ [x − b2 ][ n √ ]1/2 e− 2 (x/b) Hn (x/b)
2b ∂x 2 b n! π
1 1 1 ∂ 1 2
= [ n √ ]1/2 √ [x − b2 ] e− 2 (x/b) Hn (x/b)
2 b n! π 2b ∂x
1 1 1 2
= [ n √ ]1/2 √ e− 2 (x/b) Hn+1 (x/b)
2 b n! π 2

= n + 1ψn+1 (5.16)

The recursion relation


Hn+1 (z) = 2zHn (z) − Hn0 (z)
was used in the last step above. A similar procedure leads to

aψn = nψn−1 . (5.17)

We can abbreviate these results as the important creation and annihilation forms

a† | n > = n+1|n+1>

a|n> = n | n − 1 >, (5.18)

where ψn (x) →< x | n > . Thus a† raises the harmonic oscillator one level whereas a lowers the
harmonic oscillator one level. The sequence
√ √ √ √
a† a | n >= a† n | n − 1 >= na† | n − 1 >= n n − 1 + 1 | n − 1 >= n | n >,

or simply
a† a | n >= n | n >,
shows that
N ≡ a† a
can be associated with the quantum number n and is therefore called the number operator. In the
next section, the 1-D harmonic oscillator is described directly in terms of the N, a† and a operators

5.3.3 The Creation and Annihilation Operator description of a 1-D Os-


cillator
We can describe the 1-D oscillator in a more formal and more convenient manner by using the above
creation and annihilation operators. We arrived at these operators from their differential operator
form, but we will later generalize these procedures to more general cases for which we do not have
a differential operator representation form.
Starting from the 1-D harmonic oscillator hamiltonian
p2 1
H= + mω 2 x2 ,
2m 2
and using the definitions of N, a† and a given earlier, one can recast this Hamiltonian as simply
1 1
H = h̄ω(a† a + ) = h̄ω(N + ).
2 2
The eigenvalue equation is
H | n >= En | n >,
from which it follows with N | n >= n | n >, that
1
En = h̄ω(n + ).
2
The states are given in Dirac notation as | n >, and are realized in coordinate space by

ψn (x) =< x | n >,


CHAPTER 5. BOUND STATES 120

and in momentum space by


φn (k) =< k | n >,
with the usual connection

eikx
Z Z
ψn (x) =< x | n >= < x | k > dk < k | n >= dk √ φn (k).
−∞ 2π
The creation operator allows us to express the eigenstates in a convenient way. Starting from
the ground state | 0 >, we can generate the first excited state using

a† | 0 >= 0 + 1 | 1 >,

then the second state by repeated application


√ √ √
a† a† | 0 >= 0 + 1a† | 1 >= 0 + 1 1 + 1 | 2 > .

Carrying this out n times we have



(a† )n | 0 >= n! | n >,

or
1
| n >= √ (a† )n | 0 > .
n!
Returning to a coordinate space representation this is simply

(a† )n
< x | n >≡ ψn (x) =< x | √ |0>.
n!
The ground state | 0 > is determined by a | 0 >= 0, which is realized in coordinate space as


[x + b2 ] < x | 0 >= 0,
∂x
1 2
which has the normalized solution < x | 0 >= ψ0 (x) = 11 √1b e− 2 (x/b) . The momentum space wave
π4
functions can also be similarly determined.
The general result is:
Z ∞
(a† )n 1 1 1 d 1 2
ψn (x) = <x| √ | x0 > ψ0 (x0 ) dx0 = 1 √ (x − b2 )n e− 2 (x/b) .
n! π 4 2 n n! bn+ 12 dx
−∞

This yields the previously described wavefunctions in terms of Hermite polynomials.

Creation and Annihilation Operator Algebra


√ √
The relations: a† | n >= n + 1 | n + 1 > and a | n >= n | n − 1 >, led us to the Hermitian
† †
number operator N ≡ a a = N and

N | n >= n | n > .

Now consider the effect of aa† on any oscillator state | n >


√ √ √
aa† | n >= a n + 1 | n + 1 >= n + 1 n + 1 | n > .

Combining these results we find

(aa† − a† a) | n >= 1 | n >,

which holds for any state and hence can be described as an commutator operator relationship

(aa† − a† a) = [a, a† ] = 1 (5.19)


CHAPTER 5. BOUND STATES 121

Here 1 denotes a unit operator. As a consequence of the above basic commutator we can also deduce:

[N, a] = −a and [N, a† ] = a† (5.20)

Matrix representations of these results are: √


√0 0 0 0 0 0
   
0 1 √0 0 0 0
 1 0 0 0 0 0  0 0 2
√ √0 0 0
 
   
 0 2 0 0 0 0 0 0 0 3
√ √0 0
  
a† =  0 a=
   
 0 3 √0 0 0 

 0 0 0 0 4 √0


 0 0 0 4 0 0   0 0 0 0 0 5 
   
√ .. ..
0 0 0 0 5 . 0 0 0 0 0 .
and
 
0 0 0 0 0 0

 0 1 0 0 0 0 

 0 0 2 0 0 0 
N= .
 
 0 0 0 3 0 0 

 0 0 0 0 4 0 

..
0 0 0 0 0 .
The relation between the creation & annihilation operators and our original coordinate and
momentum operators are:
(a† + a) h̄ (a† − a)
x=b √ px = i √ .
2 b 2
From these we can obtain the expectation values
a† + a
< x >≡< n | x | n >=< n | (b √ ) | n >= 0
2
h̄ a† − a
< px >≡< n | px | n >=< n | (i √ ) | n >= 0
b 2
and the fluctuations

[a† + a] 2 h̄ 1
< x2 >≡< n | x2 | n > = √
< n | (b ) | n >= (n + )
2 mω 2

h̄ [a − a] 1
< p2x >≡< n | p2x | n > = < n | (i √ )2 | n >= h̄mω(n + ). (5.21)
b 2 2
The uncertainty principle is then satisfied: ∆x∆px = h̄(n + 1/2) ≥ h̄/2. The virial theorem result
also follows
1 1 p2
< mω 2 x2 >= En =< x > .
2 2 2m
We can also derive the off diagonal matrix element of the “dipole ” operator x, which will reappear
when we discuss dipole radiative transitions. Consider
a† + a b √ √
< n0 | x | n >=< n0 | (b √ ) | n >= √ [δn0 ,n+1 n + 1 + δn0 ,n−1 n],
2 2
hence the only nonzero matrix elements are:
r
√ h̄
< n + 1 | x | n >= n+1
2mω
and r


< n − 1 | x | n >= . n
2mω
These will be used later to explain Einstein’s stimulated and spontaneous radiation coefficients which
he introduced to derive Planck’s radiation law.
CHAPTER 5. BOUND STATES 122

5.3.4 Two and Three Dimensional Harmonic Oscillators


We can now describe a 2-D oscillator by the simple extension

p2x 1 p2y 1
H= + mωx x2 + + mωy y 2
2m 2 2m 2
to
1 1
H = h̄ωx (Nx + ) + h̄ωy (Ny + ),
2 2
where we introduce Nx = a†x ax , Ny = a†y ay , and
1 1 ∂
a†x = √ [x − b2x ],
2 bx ∂x

1 1 ∂
ax = √ [x + b2x ],
b
2 x ∂x
q

with bx = mω x
, etc.
The states are described by:

(a† )nx (a†y )ny


| nx , ny >= √x p | 0, 0 > .
nx ! ny !
This is realized as
(a† )nx (a†y )ny
< x, y | nx , ny >=< x, y | √x p | 0, 0 >
nx ! ny !
or
1 nx 1 ∂ nx ∂ ny
ψnx ,ny = ( √ ) ( √ )ny [x − b2x ] [y − b2y ] ψ0,0 .
bx 2 by 2 ∂x ∂y
Evaluation of
the derivatives bring us back to Hermite polynomials, so the
result is a simple product:
ψnx ,ny = ψnx (x) ψny (y)
1 2 1 2
with ψnx (x) = [ 2nx 1 √ ]1/2
bx nx ! π
e− 2 (x/bx ) Hnx ( bxx ), and ψny (y) = [ 2ny 1 √ ]1/2
by ny ! π
e− 2 (y/by ) Hnx y ( byy ).

Three Dimensional Case


The extension to the above discussion to 3-D is obvious. Here H = Hx + Hy + Hz , and ψnx ,ny ,nz =
ψnx ψny ψnz , with energy E = h̄ωx (nx + 1/2) + h̄ωy (ny + 1/2) + h̄ωz (nz + 1/2). For equal ω 0 s this
becomes E = h̄ω(nx + ny + nz + 3/2), which can be compared to the original 3-D case.
Another way to represent the 3-D case, with equal oscillator lengths, is to introduce vector
creation and annihilation operators ~a† , ~a,
1 i
~a† = √ (~r − p~)
2b mω
with the commutators
[ai , a†j ] = δi,j [Ni , a†j ] = δi,j a†j
or with a scalar N ≡ ~a† · ~a = Nx + Ny + Nz

[N, ~a† ] = ~a† .

The 3-D Hamiltonian is then

H = h̄ω[~a† · ~a + 3/2] = h̄ω[N + 3/2]

The relationship between the Cartesian states | nx , ny , nz > and the 3-D representation using another
set of three quantum numbers, | n, `, m > will be explored in a Homework problem. c
c The 15 operators that one can form for this problem form a representation of the SU(4) group, see later
CHAPTER 5. BOUND STATES 123

5.3.5 Hydrogenic Momentum Space Wave Functions


Observation of microscopic systems usually involve probing the system and inducing a momentum
transfer. We will see later when we discuss scattering how a momentum space wavefunction enters
into cross sections. Therefore, it is of interest to have momentum space wavefunctions.
The momentum space wavefunction has been deduced in the 1930s by several authors. The
prime reference is to Boris Podolsky & Linus Pauling, “The momentum distribution in hydrogen-
like atoms,” Phys. Rev.(2) 34 (1929), 109–116. They refer to an earlier work by H. Weyl, which
involved solving the momentum space Schrodinger equation directly. Later, V. Fock solved the
momentum space Schrodinger equation for hydrogenic atoms.
I will outline the Podolsky and Pauling method and give their result, as altered to bring it into
conformity with our definitions of the Laguerre polynomial and of the principal quantum number.
A Mathematica code for the coordinate and momentum space wave functions will be posted on
courseweb.

Spherical Bessel transform Method


To transform the 3-D coordinate wavefunction to momentum space we use
~
e−ik·~r
Z
φ(~k) = d3 r ψ(~r).
(2π)3/2
Now recall that the hydrogen wave function with definite orbital angular momentum is of the form
ψn`m (~r) = Nn` Rn` (r)Y`m (r̂)
(n−`−1)! a0 n
2
where Nn` = ( 1b )3 2n[(n+`)!] , with b ≡ 2Z , and

Rn` (r) = −e−ρ/2 ρ` L2`+1


n+` (ρ),

where ρ = r/b. To transform this state to momentum space we need the plane wave expansion
~ r ∞
e−ik·~r 2 X ∗
= (−i)` j` (kr)Y`m (r̂) Y`m (k̂).
(2π)3/2 π
`=0m

Inserting this into the above equation, we have


Z ∞ r ∞
~ 2 X 0
φn`m (k) = 2
r drdΩ (−i)` j`0 (kr)Y`∗0 m0 (r̂) Y`0 m0 (k̂) Nn` Rn` (r)Y`m (r̂)
−∞ π
`0 =0,m0
Z ∞ r
2
= r2 dr (−i)` j` (kr)Y`m (k̂)Nn` Rn` (r)
−∞ π
= Nn` <n` (k)Y`m (k̂), (5.22)
were we define the “radial” <(k) momentum space function as the spherical Bessel transform of the
radial function R(r) r Z ∞
` 2
<n` (k) ≡ (−i) r2 drj` (kr)Rn` (r).
π −∞
The requisite integral

r Z
2 3
<n` (k) ≡ −(−i) `
b ρ2 dρj` (kbρ)e−ρ/2 ρ` L2`+1
n+` (ρ),
π −∞

was shown by Podolsky & Pauling to be related to Gegenbauer polynomials Cna (z). Their result is
r s
` 2 3/2 n(n − ` − 1)! (2bk)` `+1
<n` (k) ≡ −(−i) (2b) `!4`+1 Cn−`−1 (ξ),
π (n + `)! [(2bk)2 + 1]`+2
2
−1
where ξ ≡ (2bk) a
(2bk)2 +1 and Cσ (x) is a Gegenbauer polynomial which is defined by the properties in A
& S page 794.
CHAPTER 5. BOUND STATES 124

The Momentum space radial equation for Hydrogenic Atoms


Another approach is to consider the momentum space Schrödinger equation

h̄2 2 ~
Z
k φ(k ) + d3 k 0 < ~k | V | ~k 0 > φ(~k 0 ) = E φ(~k ),

where for a Coulomb interaction
Ze2 1
< ~k | V | ~k 0 >= − .
2π 2 q 2
2
Here ~q = ~k 0 − ~k. For a bound state, we introduce E = − 2µ h̄ 2
kB and the momentum space equation
becomes Z
Z 1 1
(k 2 + kB
2
)φ(~k = d3 k 0 φ(~k 0 ).
a0 π 2 ~ ~
(k − k )0

Now we use an important relation involving both the regular (P` ) and irregular (Q` ) Legendre
functions

1 1 1 1 X
= = Q` (z)P` (cos θ),
q2 2kk 0 z − cos θ 2kk 0
`=0
0
where z = (k 2 + k 2 )/2kk 0 , and (cos θ) = k̂ 0 · k̂. The spherical harmonics can be introduced
4π X ∗
P` (cos θ) = Y`m (k̂) Y`m (k̂ 0 ),
2` + 1
`,m

so that

X
< ~k | V | ~k 0 >= < k | V` | k 0 > Y`m

(k̂) Y`m (k̂ 0 )
`=0
2
with < k | V` | k 0 >= Ze 1 4π
2π 2 2kk0 Q` (z) 2`+1 .
The final result is the momentum space radial equation

h̄2 2 2
Z
0 Ze 1 4π
k φ`m (k) + dk 0 k 2 2 Q` (z) φ`m (k 0 ) = E φ`m (k).
2m 2π 2kk 0 2` + 1
V. Fock introduced clever variables and showed that this integral equation can be solved. Note here
φ`m (k 0 ) → <`m (k 0 ).

5.3.6 Glauber States


The harmonic oscillator states can be used as a complete orthonormal basis to build wave packets.
Note that for the harmonic oscillator, all states are bound and there is no continuum; also all
states are periodic with the same period T = 2π/ω. For the 1-D case the general wave packet is a
superposition of the CON basis | n >,
X
ψ(x, t) = cn (t) | n > . (5.23)
n

Since ψ(x, t) satisfies the time-dependent Schrödinger equation with the same Hamiltonian (H =
p2 /2m + 21 mω 2 x2 ) as for the basis states, we see that cn (t) = cn e−iEn t/h̄ , and
X
ψ(x, t) = cn e−iEn t/h̄ n̄ > . (5.24)
n

One can pick any set of coefficients cn and build a wave packet. The choice made by Glauber is of
particular interest since it allows one to construct a wave packet where the phases of the various basis
states are mixed in a “coherent” manner to produce localized wave packets which have classical-like
dependence on time.
Let us examine Glauber’s choice (see the practice problems): Consider the linear combination of
states for t = 0
CHAPTER 5. BOUND STATES 125


X
| α >≡ cα
n | n >,
n=0
†n
a
where | n >= √
n!
| 0 >, are the states of a one-dimensional harmonic oscillator and the Glauber
choice is:

|α|2 αn

n ≡e
2 √ ,
n!
with α a complex number.
We can show that the state is normalized < α | α >= 1, and also that the average number of
the oscillator state in the Glauber wave packet is n̄ ≡< α | a† a | α >=| α |2 ,
In addition, we have < α | a | α >= α and < α | a† | α >= α∗ . Thus

n̄ ≡< α | a† a | α >=| α |2 =< α | a† | α > < α | a | α >,

Note that the average energy is given by


1 1
< α | H | α >= h̄ω(n̄ + ) = h̄ω(< α | a† | α > < α | a | α > + ),
2 2
which is of the classical form of an amplitude squared.
The probability of finding an oscillator in the state | n > is then a Poisson distribution
n̄n
Pn (α) = e−n̄ .
n!
It is of interest to examine the time dependence of a Glauber state. .tex

5.3.7 Normalization - Generating Functions


The normalization factor still requires discussion. For that determination and also for finding matrix
elements for other corrections and effects (such as the relativistic and spin-orbit corrections and
dipole transition), we need to develop means of evaluating matrix elements using ψ. The generating
function technique is a useful way of handling such problems. The generating function for the
Laguerre polynomials, which serves to define the polynomials used here, is

e−ρt/1−t X
a tσ−a
Ga (ρ, t) = = L σ (ρ) (−1)a (5.25)
(1 − t)a+1 Γ(σ + 1)
σ≥a

This is equivalent to our earlier definition using Laguerre’s equation, since

∂ ∂ ∂2
[t + (a + 1 − ρ) +ρ ] Ga (ρ, t) = 0
∂t ∂ρ ∂ρ2
X (−1)a 0 00
= {(σ − a)Laσ + (a + 1 − ρ)Lσa + ρ Lσa }tσ−a (5.26)
σ>a
Γ(σ + 1)
00 0
or ρ Lσa + (a + 1 − ρ)Laσ + (σ − a)Laσ = 0 as before; with σ = n + ` and a = 2` + 1! Our reason for
introducing this generating function is to enable us to find the normalization Nn` for any state. In
the appendix, we use Ga (ρ, t) to show that

2Z 3/2 (n − ` − 1)! 1 −ρ/2 ` 2`+1


ψn`m (~r) = −( ) [ ]2 e ρ Ln+` (ρ)Y`m (r̂).
a0 n 2n(n + `)!3

For example, this reduces to a simpler form for n = ` + 1 (no nodes wave function) i.e.
2Z 3/2 1 1
ψn`m (~r ) = −( ) [ ] 2 e−ρ/2 (−1)a (2n − 1)!Y`m
a0 n 2n(2n − 1)13
2Z 3/2 1 1 −ρ/2
= (+1)( ) [ ]2 e Y`m (θφ)
a0 n 2n
CHAPTER 5. BOUND STATES 126

(−1)a t0 La
using the fact that G2`+1 (ρ, 0) = 1 = σ!
σ
for σ − a = n − ` − 1 = 0,

... L2`+1
n+1 (ρ) = (−1)
2`+1
σ!

= −(2n − 1)!
Nodeless wave functions are
2Z 3 1 1 −ρ/2
ψn`m (~r ) = ( )2 ( )2 e Y`m (θφ)
a0 n 2n!
Appendix - Determination of Normalization
Normalization Nn` is determined for ψn`m = −Nn` e−ρ/2 ρ` L2`+1
n+` (ρ) Y`m (r̂)
Z Z ∞
d~r | ψn`m |2 = 1 = Nn`
2
dr r2 e−ρ ρ2` | Laσ (ρ ) |2 =
0
Z ∞
a0
2
Nn` ( n)3 dρ ρ2`+2 e−ρ | Laσ (ρ) |2
2Z 0
σ = n + `; a = 2` + 1; σ − a = n − ` − 1 > 0.
Need ∞
σ!3
Z
dρ ρa+1 e−ρ | Laσ (ρ) |2 = (2σ − a + 1)
0 (σ − a)!
2n(n + `)!3
=
(n − ` − 1)!
could prove this using generating function.
Thus
2 2Z 3 (n − ` − 1)!
Nn` =( )
a0 n 2n[(n + `)!]
Chapter 6

Scattering

References on Scattering Theory

In addition to the standard texts, Gottfried, Merzbacher, Messiah (all of which have excellent
treatments of scattering), the following books are suggested for additional study

1. L. S. Rodberg and R. Thaler “Introduction to Quantum Theory of Scattering.” On reserve in


library. My treatment is largely based on this book.
2. Wu and Ohmura “Quantum Theory of Scattering” (simpler discussion; I believe it is now
replaced by Rodberg and Thaler).

3. R. G. Newton “Scattering Theory of Waves and Particles” (excellent and detailed).


4. M. L. Goldberger and K. M. Watson “Collision Theory” (notation a problem but it is all
there).
5. Alfara and Regge “Potential Scattering” (introduction to high energy methods).

6. Mott and Massey “The Theory of Atomic Collisions” (classic book, excellent with updating).

Scattering Theory
Introduction
Much of what is known about microscopic systems is obtained by performing scattering exper-
iments. A typical set-up consists of producing a collimated beam of particles, having them strike a
target, and then detecting the number of particles scattered into a detector located in a direction
given by θ and φ.

~k 0 Detector
 :B

  B
~k  θ φ

Beam - 
Target

Figure 6.1: A scattering experiment which involves a beam a target and detector. To be a valid
experiment the setup has to assure small wave packet spreading over the dimensions of the experi-
mental apparatus.

1. The beam consists of particles such as electrons, protons, neutrons, photons, and even π ±
mesons, µ-mesons, antiprotons, K− mesons , and other exotic beams. The beam could also be atoms
of Sodium, Lithium or other such ions. Heavy ions at higher energy give us nuclear beams such as
016 nuclei. Molecular beams are also available. A wide range of energies are available; for example,
electrons can be of a few e.v. to ≥ GeV; protons up to ≥ GeV or TeV are available.
CHAPTER 6. SCATTERING 128

The characteristics of the beam, in addition to its constituent particle, is its average energy (and
deviation ∆E), the momentum, and when appropriate the state of polarization. Beams often consist
of a mixture of particles. Means for separating out particular particles, either in the beam or at the
detectors, have been developed. For polarized beams and possible detection of the polarization after
the scattering at (θ, φ), requires more elaborate double scattering set-ups.
The intensity of the beam must also be specified either in amperes or particles/sec. A high
intensity is desired to increase the counting rate and hence improve the statistics of the experiments.
Despite the need for high intensity in most cases the beams are still dilute enough so that the
scattering process can be described as a sequence of individual scatterings. That is, we assume that
each particle interferes only with itself, but the individual particles in the beam are incoherent or do
not have mutual interference. Later, we will introduce the idea of ensemble averages to describe this
property of the beam and also the state of polarization of an ensemble of particles. The microscopic
picture is:

~k
Beam  I
@ 6 ? - Target

Figure 6.2: A dilute beam with particles of various polarizations.

2. 3. For a sufficiently thin target and a dilute beam, we can analyze the above scattering process
simply as a large number of single scatterings of a beam particle from a target particle. The wave
function describing the single scattering would then give the probability amplitude for finding a
particle at a point in space. Thicker targets can be used for attenuation or absorption experiments.
One can also use gas or jet targets, emulsion techniques, etc. The subject of variations on the simple
set-up is almost the whole subject of experimental physics!
The detectors (for example, scintillation counters, or germanium detectors or spark chambers,
etc.) can be used to select out not only a specified direction but also specified energies. They can be
set to detect only certain particles, certain polarizations. For the simplest case the detected particle
is of the same type and same energy as the incident beam (elastic scattering). Of course, if the
target atoms or nuclei have structure they could absorb energy leading to inelastic scattering. The
recoil nuclei could also be detected, in general. For inelastic scattering one can also observed the
decay products of the recoiling particles. Additional counters are then needed and one can study
correlations between the observations at θ1 , φ1 and θ2 , φ2 . These correlations could also involve the
polarization states.

~k 0

: B Detector 1
   B
~k

 θ 1 φ1
Beam - X XXX
XXXθ2 φ2
Target XXX Detector 2
XX
z 


Figure 6.3: A scattering experiment which involves a beam, a target, and two detectors. The two
detectors permit coincidence experiments to measure say spin correlations.

Also reactions of various types can occur for which the exit particles differ both in nature and
number from those in the incident beam or target (production of π 0 s for example).
Having alluded to the richness of scattering experiments let me return to the simplest case of
elastic scattering of spinless particles from a structureless system. This case is represented simply
by scattering of an individual particle from a potential, which we also take to be a local, central
potential. Later we shall consider some spin dependence, noncentral and nonlocal aspects of the
CHAPTER 6. SCATTERING 129

potential. Understanding the simplest case of a central potential will serve as an introduction to the
full subject of reactions and inelastic scattering.

6.1 Stationary States for Scattering


The basic equation needed to describe the dynamics of scattering from a potential V is of course

h̄2 2 ∂
[− ∇ + V (r)]Ψ(~r, t) = ih̄ Ψ(~r, t), (6.1)
2m ∂t
for the nonrelativistic scattering of a mass m from a central,local potential V (r). The natural step
would be to form a proper wave packet description of our initial particle in the beam. a The time
development of the wave packet would then provide a complete description of the scattering process
and of the counts to be expected in the detectors.
Instead of proceeding in that manner, let us follow the procedure used for the bound state and
manufacture a set of “artificial” states which are stationary states for the continuum problem E > 0.
We consider these stationary states defined by
i
Ψ(~r, t) = e− h̄ E t ψE (~r) (6.2)

for any positive energy, with the time-independent Schrödinger equation given by

h̄2 2
[− ∇ + V (r)]ψE (~r ) = E ψE (~r). (6.3)
2m
Now solutions of this equation for E > 0 are subject to boundary conditions. Since for E > 0 there
is no exponential fall-off of ψE , we do not obtain the bound state property || ψE ||= 1 (for bound
states), but find instead that the stationary states for scattering are unbounded || ψE ||= ∞. As
before, these unbounded states can be treated using the Dirac delta function and (with the bound
states) form a complete orthonormal set of states. The complete set generated by the stationary
states (continuum plus bound states) will later be used to construct the wave packet description of
scattering. For that reason, our job is to find the solutions ψE subject to appropriate scattering
boundary conditions. We can write the stationary state Schrödinger equation in the form of a “scalar
Helmholtz equation”

2m h̄2 2
(∇2 + k 2 ) ψE (r) = V (r) ψE (r) ≡ U (r)ψE (r), E≡ k , (6.4)
h̄2 2m
which describes a scattering problem provided we introduce scattering boundary conditions. In
addition to the requirement that ψE and ∇ψE be single-valued, continuous and finite (so that ρ and
~j are reasonable), we introduce the boundary conditions for ~r → ∞ of the form

eikr
ψE (~r) −→ eikz + fk (θ, φ) . (6.5)
r →∞
~ r
Here f = amplitude of the scattered wave ≡ scattering amplitude! The above condition represents
an incident plane wave in the ẑ direction (h̄kẑ = incident momentum) and a spherical outgoing wave
as shown in the figure.
In terms of the initial plane wave’s momentum and the momentum in a direction r̂, we have with
| ~k 0 |=| ~k |= k for elastic scattering ~k = kẑ and ~k 0 = kr̂ = k k̂ 0 .
This boundary condition on ψ leads to an asymptotic flow of particles corresponding to scattering.
To understand that point, we recall that the flux density (# particles/sec–cm2 ) is given by a time-
independent expression for stationary states

j(r) = (ψ ∗ ∇ψ − ψ∇ψ ∗ ). (6.6)
2mi
a We deal with one particle scattering from V, which is described by Ψ(~ r, t) and hence can interfere with itself.
The dilute beam property assures us that it will not interfere with the subsequent scatterings of other particles. Later
we will discuss the appropriate ensemble averages needed here.
CHAPTER 6. SCATTERING 130
........................
..........
........
1
 .......
 .....
....
 ~k 0 = kr̂ = k k̂ 0
.................... ....
......... ...
...... ...
 ....
 ... ...
.............. ... ...
~k = kẑ  ....
...
...
...
....
...
...
... .. ..
- ...
.
..
.
.
..
... ... .
...
.... .
. .
.
.
. . .
............. . ...
...
.... ...
..
......
....
..... .
.......... ..
.....
......... .
......
......
........
..........
........................

Figure 6.4: An incident plane wave and an out-going spherical wave.

At large distances from the target, the wave function ψ = eikz + f eikr /r should be inserted and one
finds
~j(r) → ~jincident + ~jsc + O(1/r3 ) + interference terms (6.7)

~jincident = h̄ k ẑ = ρ ~vincident ,
m
with ρ = 1 particle /cm3 density of particles.
2
~jsc = | f | h̄k r̂.
r2 m
The O(1/r3 ) terms, which vanish for r → ∞, arise from the angle parts of

~ = r̂ ∂ + θ̂ 1 ∂ + φ̂


.
∂r r ∂θ r sin θ ∂φ
These are for example
1 ∂
(f ∗ f )θ̂ + · · ·
r3 ∂θ
(this is a tangential, not a radial flux).
The interference terms arise from combinations of the form

~ eikr
e−ikz ∇f
r
+ other terms. Note that these terms are of order 1r and r12 . Nevertheless, we drop them as is usually
done. How can this be justified? Physically, these terms are not really there since the interference
between the incident plane wave and the scattered wave is not detected because collimation of the
beam separates the plane wave from the scattered wave. A proper mathematical treatment of that
fact requires we form wave packets to describe the collimation of the plane wave. Later we will
return to justify this dropping of the interference terms. The interference that does occur for θ = 0◦
will account for the loss of flux experienced by the beam when scattering occurs.
Our procedure is to solve the scalar Helmholtz equation

(∇2 + k 2 ) ψ(r) = U (r) ψ(r) (6.8)

subject to
eikr
ψ(r) → eikz + f (r̂) .
r
for r → ∞. We define
2m
U (r) ≡ V (r).
h̄2
(+)
That will lead us to a set of stationary state solutions called ψE (r). These states, plus any bound
states, will provide a complete orthonormal (CON) basis which we will use to build wave packets,
which in turn will justify the steps we are taking now.
Cross Section - Stationary States
Using the assumption that the interference terms in ~j can be neglected and that jincident and
jscattered can be separately detected as indicated above, we can define the cross section for elastic
scattering. The set-up is
CHAPTER 6. SCATTERING 131

~r dA Detector
:B

  B
~k  θ φ
Beam - 
Target

dA
Figure 6.5: A beam strikes a target and hit a detector of area dA and of solid angle r2 .

The number of particles/sec seen by the detector of area dA is

~ =| f |2 h̄k dA =| f |2 vi dΩ.
~jsc · dA
m r2
The scattering cross-section is defined by
particles ~
scattered into dA at θ, φ ~jsc · dA
dσ = sec
= h̄k
=| fk (θ, φ) |2 dΩ
incident flux density m

particles
sec
dσ = # particles = cm2 .
sec−cm2

The differential cross section is therefore given by



=| fk (θ, φ) |2 (6.9)
dΩ
for elastic scattering. The total elastic scattering cross section is given by
#particles
into all angles!
Z Z
dσ sec
σtot = dΩ = dσ = #particles
.
dΩ sec−cm2

Later, when we consider nonhermitian potentials and the possibility of inelastic collisions the defi-
nition of the total cross section will be extended.
Cross sections have the units of an area. Typical units used are:
1 barn 10−24 cm2 = 100F 2 ; 1F=1 Fermis = 10−13 cm.

mb 10−27 cm2 = 1
10 F
2

µb 10−30 cm2 = 10−4 F 2


nb 10−33 cm2 = 10−7 F 2

6.1.1 Integral Equation For Scattering


7. For the stationary state scattering solution, we have a well-defined mathematical problem; namely,
to solve the scalar Helmholtz equation

(∇2 + k 2 )ψE (~r) = U (r)ψE (~r) (6.10)

2m h̄2 2
U (r) =
2 V (r) ; E = 2m k ,

subject to the asymptotic boundary condition with an incoming plane wave and an outgoing spherical
wave:
1 i~
k·~
r eikr
ψE (~r ) −→ (e + fk (θ, φ) ).
r→∞ (2π)3/2 r
CHAPTER 6. SCATTERING 132

Here we set ~k = kẑ and introduce (2π)−3/2 for convenience in the later formal scattering theory
1 3
version. A normalization of ρ = (2π) 3 particles /cm (or 1/vol for box normalization case.) The
dσ 2
expression dΩ =| f | is unchanged!
We can proceed under the assumption that the above boundary condition does indeed describe
asymptotic states consisting of the beam and scattered waves. As we have seen that requires further
justification (to drop interference terms in the current ~j) which we will discuss later by invoking the
full time dependent wave packet description.
First let us solve the above well-defined mathematical problem so as to obtain a useful complete
orthonormal CON set of states. The differential equation plus boundary conditions can be rewritten
as an integral equation in which the asymptotic boundary condition is incorporated; we have
~
eik·~r
Z
ψE (~r ) = + (~r | G(k 2 ) | ~r 0 ) U (~r 0 ) ψE (~r 0 ) d~r 0 . (6.11)
(2π)3/2

Let us check this is the form of the solution by applying (∇2 + k 2 ) to both sides.
The Green’s function satisfies b

(∇2r + k 2 )(~r | G(k 2 ) | ~r 0 ) = δ(~r − ~r 0 ) (6.12)

to assure that (∇2 + k 2 )ψE (r) = U (r)ψE (r). We must now design a solution to this differential
equation incorporating the spherical outgoing wave f eikr /r. To find that solution, it is simplest to
Fourier transform this equation; i.e. to use a momentum space representation. Since (~r | G(k 2 ) | ~r 0 )
is a function of (~r − ~r 0 ) (because the delta function is), we can write.
~ 0 ·(~ r 0)
r −~
e ik
Z
2 0
(~r | G(k ) | ~r ) = Gk2 (~k 0 ) d~k 0 .
(2π)3

Inserting this into (∇2 + k 2 )G = δ(r − r 0 ) gives us

(k 2 − k 02 )Gk2 (~k 0 ) = 1

or
1
Gk2 (~k 0 ) = .
k 2 − k 02
This last step is not well defined when k = k 0 . To make it well defined, and to properly incorporate
the boundary conditions, we introduce the complex k 0 –plane and move the k 0 = ±k singularities off
the real axis to remove the singularity. Thus, we have the Green’s function defined by
~ 0 ·(~ r 0)
eik r −~
d~k 0
Z
2 0
(~r | G(k + i) | ~r ) = , (6.13)
(2π)3 k 2 − k 02 + i

where it is understood that one takes the limit  → 0 after evaluation of the integral. The role of 
is to move the singularity off the real axis. In the complex k 0 –plane the singularity is located at
r
02 2 0 i
k = k + i or k = ± | k | 1 + 2
k
i
k0 = ± | k | ±
2|k|
located in the figure at points marked by a × and by a dot.
0
−~
We can now evaluate the integral, using the Residue, or Cauchy, theorem since eik(~r r)
is an
analytic function. We use
F (z 0 ) 0
I
dz = 2πi F (z),
z0 − z
where F (z) is analytic inside a closed contour.
b (∇2 + k2 )eik·r = 0
CHAPTER 6. SCATTERING 133

×
·

(+)
Figure 6.6: The complex k 0 plane with the singularities in the out-going wave Green’s function G0 .
The k 0 = k + i 2|k|

pole is marked by a × and the k 0 = k − i 2|k|

pole by a dot.

Some preparatory steps are needed. With d3 k = k 2 dkd cos θdφ = k 2 dkdµdφ, we have
∞ +1 0
dk 0 k 02 2π dµ eik Rµ
Z Z
(~r | G(k 2 + i) | ~r 0 ) = − 02
0 k − k 2 − i −1 (2π)3
∞ 0 0
dk 0 k 0 2π (eik R − e−ik R )
Z
=−
0 (k 02 − k 2 − i) iR (2π)3
Z ∞ 0
1 dk 0 k 0 eik R 1
=− 2 02 2
( ).
(2π) −∞ (k − k − i) iR
Here R ≡| ~r 0 − ~r | . The contour can be described as shown below with the contribution of the
large semi-circle vanishing as k 0 → i∞.
...........................................
.....
............ .. .. ..........
......... ........
.
......... .....
.. .....
....
....
...  ........
.. i∞ ...
... ...
.. ...
.... ...
...
..
... ...
...
.... . ×
- .. . - .
−∞ ∞

Figure 6.7: The complex k 0 plane with the contour closed in the upper half complex plane, thereby
including the pole or singularity at k 0 = k + i 2|k|

pole as marked by a ×. The infinite upper circular
contour’s radius is denoted by i∞.

Therefore
1 1 k ikR 1 eikR
(~r | G(k 2 + i) | ~r 0 ) = − 2πi e = −
(2π)2 iR 2k 4π R
0
2 0 (+) 1 eik|~r−~r | 0
(~r | G(k + i) | ~r ) ≡ (~r | G (E) | ~r ) = −
4π | ~r − ~r 0 |
The conclusion is that ψE (r) is given by
~
eik·~r
Z
(+) (+)
ψE (~r ) ≡ ψ~ (~r ) = + (~r | G(+) (E) | ~r 0 ) d~r 0 U (r0 ) ψ~ (~r 0 ). (6.14)
k (2π)3/2 k

The notation has been extended to include the energy and the direction of the incident beam by
h̄2 2
using the label ~k instead of E = 2m k . Also the outgoing spherical wave boundary condition has
been indicated by the (+) superscript. Solutions of this type are also called out solutions. To
show that this integral equation has properly incorporated the outgoing spherical wave boundary
condition, consider the definition of R =| ~r − ~r 0 |, where ~r 0 is confined to the range of V (r0 ). We
assume here that V (r0 ) is of finite range:
p r0 r0
R= r2 + r02 − 2rr0 cos α = r{1 + ( )2 − 2 cos α}1/2 ;
r r
note that cos α = r̂ · r̂ 0 .
CHAPTER 6. SCATTERING 134

As ~r → ∞ and since inside the integral r0 ≤ the range of V (r0 )

r0
R −→ r{1 − cos α} = r − r r0 cos α = r − r̂ · ~r 0
r
and
1 eikr
Z Z
(+) ~ 0 ·~ 0
lim (~r | G(+) (k 2 ) | ~r 0 ) U (r0 ) d~r 0 ψ~ (~r 0 ) = − e−ik r
U (r0 ) ψ~k (r 0 ) d~r 0 .
r→∞ k 4π r

The relation ~k 0 = k r̂ has been used to change kr̂ · ~r 0 to ~k 0 · ~r 0 . Thus we get


~
eik·~r 1 eikr
Z
(+) ~ 0 ·~ 0
ψ~ (~r) −→ 3/2
− e−ik r
U (r0 ) ψ~k (~r 0 ) d~r 0
k r→∞ (2π) 4π r

1 ikz eikr
−→ {e + f k (θ, φ) }
(2π)3/2 r
with the scattering amplitude given by
~ 0 0
e−ik ·~r
Z
2m (+)
fk (θ, φ) ≡ f (+) (~k 0 , ~k) = −2π 2 ( 2 ) V (r0 ) d~r 0 ψ~ (~r 0 )
h̄ (2π)3/2 k

2m (+)
) < ~k 0 | V | ψ~ >
= −2π 2 (
h̄2 k

(+)
= −(4π 2 m/h̄2 ) < ~k 0 | V | ψ~ > . (6.15)
k

Since V is a function of | ~r 0 | only, the scattering amplitude depends on k̂ 0 · k̂ = cos θ and k only,
not on the azimuthal angle φ.

Incoming Spherical Wave Solution


Another boundary condition for which there are stationary state solutions is provided by considering
incoming spherical waves e−ikr /r. These states provide an alternate complete set (with the bound
states) and prove to be very useful later. These incoming, spherical waves are generated by picking
a different Green’s function defined by
~ 0 ·(~ r 0)
eik r −~
d~k 0
Z
(~r | G(k 2 − i) | ~r 0 ) =
(2π)3 k 2 − k 02 − i

The singularities now occur at k 0 = ± | k | ∓ 2|k|


i
and the contour is

...........................................
.........
........... ..........
..
...........
. ........
......
..
... .....
..
.. ....
...
. ....
... ...
... ...
... ...
.... ...
... ...
.. ...
..... ×
...
...
. - . -
−∞ · ∞

Figure 6.8: The complex k 0 plane with the contour again closed in the upper half complex plane,
but now including the pole or singularity at k 0 = −k − i 2|k|

pole as marked by a ×.

We again evaluate the integral by taking the limit R → ∞ and now closing the contour down.
Z ∞ 0
1 dk 0 k 02 eik R 1 2π 1 (−k 2 ) ikR
(~r | G(k 2 − i) | ~r 0 ) = − 2 02 2
=− e
(2π) −∞ k − k + i iR (2π) iR (−2k)
CHAPTER 6. SCATTERING 135

0
1 e−ikR 1 e−ik|~r−~r |
=− =− ≡ (~r | G(E − i) | ~r 0 ) ≡ (~r | G(−) (E) | ~r 0 ) = (~r | G(+) (E) | ~r 0 )∗
4π R 4π | ~r − ~r 0 |
Now the residue at k 0 = − | k | + 2|k|
i
is picked up. The solution is therefore

~
eik·~r
Z
(−) (−)
ψ~ (~r) = + (~r | G(E − i) | ~r 0 ) d~r 0 U (r0 ) ψ~ (~r 0 ).
k (2π)3/2 k

This of course corresponds to incoming spherical waves plus an “outgoing” plane wave.

The corresponding scattering amplitude (or amplitude of the incoming spherical wave) is deter-
mined to be
Z +i~k 0 ·~r 0
4π 2 m (−) 4π 2 m e (−)
f (−) (~k 0 , ~k) = − 2 < −~k 0 | V | ψk >= − 2 V (r) ψk (~r 0 ) d~r 0 . (6.16)
h̄ h̄ (2π)3/2

Another comment, which is related to reversal of motion (time reversed states), is that ψ (+) and
(+)∗ (−)
ψ (−) are related by complex conjugation plus reversal of ~k : ψ ~ (~r) = ψ~ (~r).
−k k

Standing Wave Stationary State Solutions


A third boundary condition often used for stationary states, is obtained by combining the previous
out and in cases. Since this case yields no radial flux other than the plane wave, it is called a
standing wave boundary condition. It is determined by
1 1 cos kR
(~r | G(E + i) + G(E − i) | ~r 0 ) = − ≡ (~r | G(0) (E) | ~r 0 )
2 4π R
∞ ~ 0 0 Z ∞ ~ 0 i~k 0 ·R~
eik ·(~r−~r ) (k 2 − k 02 )
Z
1 1 1 dk e
= { + } = ,
2 −∞ (2π)3 k 2 − k 02 + i k 2 − k 02 − i −∞ (2π)3 (k 2 − k 02 )2 + 2
c
∞ ~ 0 ~
d~k 0 eik ·R
Z
=P
−∞ (k 2 − k 02 )(2π)3
etc.

6.1.2 Formal Version of Stationary State Scattering


The previous explicit construction of stationary state solutions to the scalar Helmhotz equation can
be written in the abbreviated, representation-independent manner of Dirac notation. Again I will
show how to use Dirac notation by example and will refer you to the standard sources (especially
Dirac) for more information.
Our stationary state wave equation is formally

(H0 + V ) | ψE >= E | ψE >, (6.17)


c Using HW problem we have
1 x − x0
P = lim
x − x0 →0 (x − x0 )2 + 2
or
1 P
= ∓ πiδ(x − x0 )
x − x0 ± i x − x0
1 1 1 1
[ + ]=P
2 x − x0 + i x − x0 − i x − x0
CHAPTER 6. SCATTERING 136

with H0 the “unperturbed Hamiltonian” being the kinetic energy operator H0 = p2 /2m. Here we
deal with operators and states in a Hilbert space. The stationary states are, as before, defined by
i i
| ΨE (t) >= e− h̄ Et | ψE >= e− h̄ Ht | ψE (0) > . (6.18)

We can rewrite the above as


(E − H0 ) | ψE >= V | ψE > . (6.19)
For E > 0, continuum solutions to the homogeneous equation exist (but not for bound states). These
solutions are simply the incident plane wave state (E − H0 ) | ~k >= 0, which are also eigenstates of
h̄2 2
p~ | ~k >= h̄~k | ~k > with ~k = kẑ; E = 2m k . Adding these solutions to the homogenous equations,
we have
1
| ψE >=| ψ~k >=| ~k > + V | ψ~k > . (6.20)
E − H0
h̄2 /2m
Calling G0 (E) = E−H0 , we have various solutions

(±)
| ψ~ >=| ~k > +G0 (E ± i) U | ψ~± >
k k

2m
U≡ V
h̄2
h̄2 /2m
G0 (E) ≡ .
E − H0

The ~r - Representation
To see that we have properly generated our original explicit realization of these abstract equations,
let us consider the Green’s function operator

h̄2 1 X | n >< n | h̄2



0 = G0 (E ± i) = =
2m E ± i − H0 n
E + i − En 2m

The states | n > are eigenstates (a complete orthonormal set) of the kinetic energy operator H0 . Of
course, these are simply | n >→| ~k > since

h̄2 02 0
H0 | k 0 >= k |k >.
2m
Correspondingly, the sum over n becomes an integral (as in the case of our box normalization
discussion) and we return to
~ 0 0
e−ik ·(~r −~r )
Z
(~r | G0 (E ± i) | ~r 0 ) = d~k 0
(2π)3 (k 2 ± i − k 02 )
our old case. Therefore
h̄2 /2m 1 e±ikR
(~r | | ~r 0 ) = −
E ± i − H0 4π R
with R =| ~r − ~r 0 | . Note we have introduced the subscript on G0 to record the appearance of H0
in (E − H0 )−1 ; we will also deal with the exact Green’s function (for stationary states) defined by
G(E) → (E − H)−1 with the exact Hamiltonian H.
Taking the r-representation of our formal equation | ψ >=| k > +G0 U | ψ >, we return to our
original integral equation

h̄2 /2m
Z
(±) (±) (±)
< ~r | ψ~ >= ψ~ (~r ) =< ~r | ~k > + (~r | | ~r 0 )(~r 0 | U | ψk > d3 r0 .
k k E ± i − H0
For a general nonlocal potential (useful later)
Z
(±) (±) (±)
(r | U | ψk >= (r0 | U | r00 )ψk (r00 )d3 r00
0
→ U (r0 )ψ~ (r0 ).
local case k
CHAPTER 6. SCATTERING 137

A local potential has the property

(r0 | U | r00 ) = δ(~r 0 − ~r 00 )U (r0 ).

Thus we get
~
eik·~r eikR
Z
(±) 1 0 (±)
ψ~ (~r ) = − d~r U (r0 )ψ~ (r0 ).
k (2π)3/2 4π R k

Momentum Space Representation


The formal equation
(±) 1 (±)
| ψ~ >=| ~k > + V | ψ~ > (6.21)
k E ± i − H0 k

has the advantage of being representation-independent. In the previous section, we took the r-
representation and found that the original integral equation for scattering was obtained. Now we
take the momentum space representation.
First recall that ~k refers to the energy E = h̄2 k 2 /2m which is fixed by the beam; the direction
of the beam is ~k = kẑ. The momentum space representation involves taking
(±) (±)
< ~k 0 | ψ~ >= ψ~ (~k 0 )
k k

0
where k refers to the variable momentum coordinate; it is not the momentum of the scattered wave.
With this in mind, the equation for scattering expressed in momentum space is

(±) h̄2 /2m (±)


ψ~ (~k 0 ) = δ(~k − ~k 0 ) + (~k 0 | U | ψ~ >
k E ± i − H0 k

(±)
(~k 0 | U | ψ~ >
= δ(~k − ~k 0 ) + k
. (6.22)
k 2 ± i − k 02
This follows immediately from the Schrödinger equation in momentum space since (H0 + V ) | ψ >=
E | ψ > is
h̄2 2
(k − k 02 )ψ(k 0 ) = (k 0 | V | ψ)
2m
(k 2 − k 02 )ψ(k 0 ) = (k 0 | U | ψ). (6.23)
For plane wave solutions (k 2 − k 02 )δ(~k − ~k 0 ) = 0 we get the above result. For bound states we get
h̄2 2
k = ikB , E = − 2m kB and
(k 0 | V | ψB )
ψB (k 0 ) = 2 . (6.24)
k 2 + kB
(±) R ~0 (±)
Note that (~k 0 | U | ψ~ >∼ eik ·~r U (r) ψ~ (~r ) d~r is not directly the scattering amplitude
k k
since
4π 2 m (±)
f (±) (~k 0 , ~k) = − 2 < ±~k 0 | V | ψ~ >
h̄ k

(±)
= −2π 2 < ±~k 0 | U | ψ >| ~ ~ 0
~ |k|=|k |
k

is evaluated with | ~k |=| ~k |= k. We evaluate the above expression on the energy shell to get the
0
(±)
scattering amplitude. However, < ~k 0 | U | ψ~ > as it appears in the equation
k

(±)
(±)
(~k 0 | U | ψ~ )
ψ~ (~k 0 ) = δ(~k − ~k 0 ) + k
k k 2 ± i − k 02

has a variable ~k 0 which could be off-the-energy-shell: | ~k 0 |6=| ~k | .


CHAPTER 6. SCATTERING 138

Born Approximation
The Born approximation in lowest order consists of taking

ψ '| k >,

inside the range of potential as it appears in the expression for the scattering amplitude

4π 2 m ~ 0 4π 2 m
fBorn = − 2 < k | V | ~k >= − 2 V(~q).
h̄ h̄
The lowest order Born approximation is applicable at high energies and/or for weak potentials away
from resonances and bound state effects- see later for the criteria for the use and convergence of
the Born approximation series based on the product of the potential strength times the finite range
squared.
For the Coulomb interaction, one might use:

Ze2
V =− V(q) =
r
and
π 2 m2
| V(q) |2
σBA (θ) = 16
h̄2
to arrive at the Rutherford scattering formula, but it is important to realize the infinite-range of
the Coulomb interaction does not allow us to use the Born approximation blindly; in addition our
scattering amplitude result was obtained under the assumption of a finite ranged potential. It is an
important coincidence that the full treatment of the Coulomb scattering problem differs from the
lowest order result by a complicated phase.

6.1.3 The Transition Matrix and the Lippmann-Schwinger Equation


(+)
The quantity (~k 0 | V | ψ~ > (which is a function of ~k 0 and ~k for all values of ~k and ~k 0 ) is called
k
the transition or T-matrix when written as:
(+)
(~k 0 | V | ψ~ >≡ (~k 0 | T (E + i) | ~k); (6.25)
k

T is an operator defined between plane wave states. We can derive an integral equation for this
quantity, which is simply another version of the integral equation for scattering. Using

(+) 1 (+)
| ψ~ >=| ~k > + V | ψk > (6.26)
k E + i − H0
we get
1 (+)
(~k 0 | T (E + i) | ~k) = (~k 0 | V | ~k) + (~k 0 | V V | ψ~ >
E + i − H0 k
Z
1 (+)
= (~k 0 | V | ~k) + (~k 0 | V | ~k 00 ) h̄2
(~k 00
| V | ψ~ > d3 k 00
2 i0 − k 002 ) k
2m (k +

(~k 0 | V | ~k 00 )(~k 00
| T (E + i) | ~k)
Z
= (~k 0 | V | ~k) + h̄2
d3 k 00 .
2m (k
2 + i0 − k 002 )
This can be written formally as
1
T (E + i) = V + V T (E + i) (6.27)
E + i − H0
–called the Lippmann-Schwinger equation.
The scattering amplitude is obtained from the on-shell T-matrix using

4π 2 m (+) 4π 2 m
f (+) (~k 0 , ~k) = − 2 < ~k 0 | V | ψ~ >= − 2 < ~k 0 | T (E + i) | ~k >||~k|=|~k 0| . (6.28)
h̄ k h̄
CHAPTER 6. SCATTERING 139

Knowledge of T for all values of ~k and ~k 0 , permits one to calculate the full wave function (for all
values of ~r or in momentum space for all ~k 0 using

(+) (2m/h̄2 )
ψ~ (~k 0 ) = δ(~k 0 − ~k) + 02
< ~k 0 | T (E + i) | ~k > . (6.29)
k k2 + i − k
Finally, we can write this in r-space as
~ 0 ~ ~ 0 ·~
eik ·~r ~ 0 (+) ~ 0 eik·~r e ik r
< ~k 0 | T (E + i) | ~k >
Z Z
(+)
ψ~ (~r ) = 3/2
dk ψ~ (k ) = + d~k 0
h̄2
, (6.30)
k (2π) k (2π)3/2 2m (k
2 + i − k 02 )

or
(+) 1 (+)
| ψk >=| ~k > + V | ψ~ >
E + i − H0 k

1
=| ~k > + T | ~k > . (6.31)
E + i − H0

Formal Application
As examples of the formal version of the integral equation for scattering, let us again consider the
Hilbert space rendition. We start with the stationary state equation

(H0 + V ) | ψ >= E | ψ >,


2
p h̄2 2
where E > 0 and H0 denotes the kinetic energy operator 2m and we define E = 2m k0 . We recast
this as
(E − H0 ) | ψ >= V | ψ >,
and note that the plane wave solution holds for H0 alone

(E − H0 ) | ~k0 >= 0.

The next step in the formal version is

(+) 1 (+) 1
| ψ~ >=| ~k0 > + V | ψ~ >=| ~k0 > + T | ~k0 >,
k0 (E + i − H0 ) k0 (E + i − H0 )

where we have invoked the definition of the transition operator


(+)
T | ~k0 >= V | ψ~ > .
k0

The notation includes the incident plane wave vector ~k0 and the out-going wave boundary condition
labelled by the superscript (+).

The Møller Operator


The last result can also be written as

(+) 1
| ψ~ >= (1 + T ) | ~k0 >= Ω(+) (E) | ~k0 >,
k0 (E + i − H0 )
where
(+)
Ω(+) (E) ≡ 1 + G0 (E)T (+) (E)
is called the Møller or distortion operator. The meaning of Ω(+) (E) = Ω(E + i) is that it constructs
(+) (+) (+)
ψ~ from the plane wave state; also, we have < ~k 0 | Ω(+) (E) | ~k0 >=< ~k 0 | ψ~ >= ψ~ (~k 0 ). The
k0 k0 k0
momentum space matrix element of the Møller is thus the scattering wave function in momentum
space. It is later useful in defining the S-matrix (or scattering matrix).
CHAPTER 6. SCATTERING 140

The Low Equation


Another application of these formal scattering theory manipulations is to derive the Low equation.
Here I will go all the way and present the derivation in full formal style, with the understanding
1
that every step can be made explicit, if needed. The derivation uses Ω(ω) = 1 + ω−H 0
T (ω), where
(±)
ω is a possibly complex energy variable ω = E ± i, with Ω (E) = Ω(E ± i). Using the notation
| φ >=| ~k0 >, and G0 (ω) = (ω − H0 )−1 , we have

| ψ >=| φ > +G0 V | ψ >

or
(1 − G0 V ) | ψ >=| φ >
| ψ >= (1 − G0 V )−1 | φ >= Ω | φ > .
Thus
Ω = (1 − G0 V )−1 ,
and
Ω−1 = (1 − G0 V ) = G0 (G−1
0 − V ) = G0 (ω − H0 − V ) = G0 (ω − H).

Another equivalent way to get this result is

| φ >= (1−G0 V ) | ψ >= G0 (G−1


0 −V ) | ψ >= G0 (ω −H0 −V ) | ψ >= G0 (ω −H) | ψ >= Ω
−1
| ψ >,

and again
Ω−1 = G0 (ω − H).
Now
1 1 1
Ω= G−1 = (ω − H0 − V + V ) = 1 + V.
ω−H 0 ω−H ω−H
Here we use the inverse operator property (AB)−1 = B −1 A−1 .
Since V | ψ >= V Ω(ω) | φ >= T | φ >, we have T (ω) = V Ω(ω), and we arrive at the Low
equation
1
T (ω) = V + V V.
ω−H
1
The exact Green’s function ω−H enters here, which is the main feature of the Low equation. However
this is not a solution equation since to determine the solution we already need the solution , i.e. we
need to know the exact Green’s function.
Nevertheless, the Low equation is useful; for example by introducing the complete set of exact
scattering plus bound states in closure form we see that
X V | n >< n | V
T (ω) = V + ,
n
ω − En

which displays bound state poles at ω = − | EnB |, and a continuum of singularities ( a branch cut)
for all real positive ω values ω = E > 0.
The discontinuity across this positive energy E > 0 continuum branch cut is then d
X 1 1
T (E + i) − T (E − i) = V | n >< n | V [ − ]
n
E + i − En E + −i − En
X
= −2πi V | n > δ(E − En ) < n | V δ(E − En ).
n
d The 1 P
relations E±i−H
= E−H
∓ πiδ(E − H) based on contour integrations (see earlier HW) was invoked here.
CHAPTER 6. SCATTERING 141

Formal version of the Optical theorem


We can make X
T (E + i) − T (E − i) = 2πi V | n > δ(E − En ) < n | V
n

more explicit by (1) taking diagonal matrix elements in momentum space and (2) using | n >→|
(+)
ψ~ 0 >, where we can omit bound states because the delta function δ(E+ | EBn |) vanishes for
k
E > 0. We have then
Z
(+) (+)
< ~k | T (E + i | ~k > − < ~k | T (E + i | ~k >∗ = −2πi < ~k | V | ψ~ 0 > δ(E − Ek0 ) < ψ~ 0 | V | ~k >,
k k

We relate this to the scattering amplitudes using


m
f (+) (~k | ~k) = −(2π)2 2 < ~k | T (E + i | ~k >,

where we have taken ~k 0 → ~k, e.g. forward angle scattering is examined. We also use T † (E − i) =
T (E + i) to identify the second term as f ∗(+) (~k | ~k). With these factors, we obtain

h̄2 2 1 dk 0
Z Z
m
f (+) (~k | ~k)−f ∗(+) (~k | ~k) = (2π)2 (2πi)( ) | f (+) (~k 0 | ~k) |2 dΩk̂ 0 k 02 dEk0 δ(E−Ek0 ),
h̄2 m (2π)4 dEk0
dEk0 h̄2 0
with dk0 = mk , we get
Z
i

2iImf (0 ) = | f (+) (~k 0 | ~k) |2 dΩk̂ 0

and finally the optical theorem

Imf (0◦ )
σtotal =
k
This is a formal version of the optical theorem, which we will now examine in several other, less
formal, ways.
CHAPTER 6. SCATTERING 142

6.1.4 The Optical Theorem - Stationary State Version


Before discussing the wave packet description of scattering, I would like to present a version of the
(+)
“optical theorem” using our stationary states ψk . Later we shall rederive this result in a more
rigorous fashion using wave packets, but I think this theorem is so important that we should see
both versions. (I base the discussion on the one given by Rodberg and Thaler.)
Earlier we found that the flux density included interference terms
~j = jsc + jincident + O(1/r3 )+ interference terms between eikz and f eikr /r. We now examine these
interference terms and subject them to the condition that we have no flux lost and hence ∇ ~ · ~j =

− ∂t ρ = 0 for stationary state scattering with no absorption. Thus we require
Z
~=0
~j · dS

ikr
over a sphere at large r. Let us evaluate ~j again using ψ → (eikz + f e r ) (2π)13/2 for large r.

h̄ h̄k | f |2
j→ (ψ ∗ ∇ψ − ψ∇ψ ∗ ) = ρ{ẑ + 2 r̂ + interference terms } + O(1/r3 )
2im m r
ρ = 1/(2π)3
interference terms are
1 −ikz eikr e−ikr
= [e r̂ f + eikz f ∗ r̂ ]
2 r r
1 e−ikr ikz eikr
+ [f ∗ ẑe + f ẑ e−ikz ] + O(1/r2 ).
2 r r
Note we drop O(1/r2 ) terms in the interference current, which really needs the justification of our
later wave packet discussion.

~jinterf = 1 (ẑ + r̂)Re[fk (θ, φ)eik(r−z) ]


r
~ = r̂ dΩ r2 in the flux conservation ~j · dS ~ = 0 condition, we get
R
Using dS

r2
Z Z Z
h̄k
0= ρ{ | f |2 dΩ + dΩ ẑ · r̂ r2 + dΩ (r̂ · ẑ + r̂ · r̂) Re (f eik(r−z) }.
m r
Now recall Z Z
2 dσ
| f | dΩ = σT OT = dΩ
dΩ

(Note, we have a logical inconsistency here since the result dΩ =| f |2 so far has been “derived”
under the assumption that jinterf = 0.) Such puzzles will be cleared up later. On with the usual
stationary state rendition.
Now, we have Z Z 1
dµ dφ cos θ r2 = dµ µ · · · = 0
−1

Thus with µ = cos θ Z


σtotal = −Re dφ dµ r(1 + cos θ)f (θ, φ)eikr(1−µ)
Z
= −Re r dφ dµ (1 + µ) f (θ, φ)eikr(1−µ)

Now we can now show that for large r


Z
4πi
r dφ dµ (1 + µ) f (θ, φ)eikr(1−µ) → f (θ = 0◦ ) + O(1/r2 ) + · · ·
k
Hence
2i 4π
σtotal = 2πRe( f (θ = 0◦ )) = Im( f (θ = 0◦ )).
k k
CHAPTER 6. SCATTERING 143

The result

σtotal = Im( f (θ = 0◦ ) ).
k
is called the “optical theorem” or the “shadow theorem.” It was originally derived for Quantum
Mechanical scattering by Bohr,Peierls and Placzek. It states that the imaginary part of the forward
scattering amplitude determines the total cross section (i.e. everything that can occur at all angles).
We have only discussed elastic scattering but the generalization to inelastic collisions and reactions
holds true. Basically the forward angle is where the shadow forms by destructive interference; then
by flux conservation the missing flux goes elsewhere, into other processes.
The physical meaning is again that the plane wave + scattered wave interfere in forward direction
and deplete forward flux thereby producing a shadow (lower intensity at zero degrees); the missing
flux goes into other scatterings and reactions.
Could do the derivation for a nonhermitian, absorptive potential V + 6= V and we find that the
theorem holds more generally than we have reasoned. The total cross section can be divided into
scattering and reaction terms
σtotal = σSC + σreactions .
See later.
CHAPTER 6. SCATTERING 144

PHYSICS 266 – SPRING, 2005


LECTURE SCHEDULE
29 LECTURES IN 14 WEEKS

The following table lays out the lecture plans.


This schedule will no doubt be revised as the course proceeds.

NOS. DATE TOPIC


1. Jan. 6 TH Wave Packet Theory of Scattering
2. Jan. 11 T Cross Section and Optical Theorem
3. Jan. 13 TH Partial Wave Expansion - Radial Integral Equation
Lippmann Schwinger Equation- Phase Shifts, Cross Sections
4. Jan. 18 T Phase Shifts, Cross Sections
5. Jan. 20 TH T-Matrix - Bound States - Resonances - Thresholds
6. Jan. 25 T Stern-Gerlach Experiment Spin-Magnetic Moment
Electron spin, Operators, States, and Pauli Matrices
7. Jan. 27 TH Two Spins, Orthonormality - Spin in Scattering
8. Feb. 1 T Properties of State Vectors Operators
Adjoint, Hermitian, Unitary
9. Feb. 3 TH Unitary Transformations - Symmetries Translation
10. Feb. 8 T Rotation, Invariance, Scalar and Vector Operators
11. Feb. 10 TH Addition of Angular Momentum Define CG, 3-j and 9-j Symbols
Euler Angles and Wigner D functions
12. Feb. 15 T Tensor Operators - Wigner-Eckart Theorem
13. Feb. 17 TH Quantum Dynamics - Perturbation Theory Secular Equation
14. Feb. 22 T Brilloun-Wigner Perturbation Theory
15. Feb. 24 TH Rayleigh Schrodinger Perturbation Theory
16. March 1 T Degenerate Perturbation Theory
17. March 3 TH Time Dependent Perturbation Theory-Golden Rule
Selection rules and intensities
18. March 15 T Many Body problem and Pauli Principle
19. March 17 TH Many Body problem and Pauli Principle
20. March 22 T Applications of Perturbation Theory - Hydrogen Fine Structure
Proton Size-Zeeman Effect - Lamb Shift
21. March. 24 TH Quantized Electromagnetic Field
22. March 29 T Second Quantization Uncertainty principle for fields
23. March 31 TH Quantized Electromagnetic Field
24. April 5 T Selection Rules and Intensities
25. April. 7 TH Black-Body Radiation
26. April 12 T Dirac Equation (Time permitting)
27. April 14 TH Dirac and Hydrogen
28. April 19 T Dirac Equation and Field Theory
29. April 21 TH Dirac Equation and Field Theory
CHAPTER 6. SCATTERING 145

Welcome back. Before starting our second half, I would like to discuss some questions of policy
and approach. One reason for repeating this step is to make our procedures clear and another reason
is to solicit your opinions concerning the best procedures.

Policies

Concerning our texts, we shall again use Vols. I & II of Messiah. Several others books will be on
reserve. I suggest you familiarize yourselves with a variety of books and references. The content of
the course is defined by the lectures and homework. I will again put my lecture notes and homework
solutions on courseweb. I urge you to formulate your own set of notes and go over the derivations,
which is a very good way of learning.
I will schedule two exams plus a final. The first two exams count 20% each, the final counts 35%
and your homework grade counts 25%.
Tentative Exam Dates:
Exam # 1 Sat. Feb 12th.
Exam # 2 Sat. April 2 th.
FINAL EXAM During week of April 25
One of these might be a take-home, depending on how you (and I) feel about that type of exam.
I will schedule an optional recitation on Wednesdays 4-5.
Chapter 7

Scattering-Continued

7.1 Wave Packet Description of Scattering


Reference: Messiah pages 372-380 and Rodberg & Thaler
Last term we saw that we could “prove” the optical theorem using stationary state scattering
states and the requirement of flux conservation. That proof, and the discussion of how to define
the cross section, was somewhat marred by the manner in which some of the interference terms of
order 1/r2 were treated. To fully justify the steps taken during the stationary state discussion, we
now turn to the full time-dependent, wave-packet discussion. In addition to justifying our earlier
steps, the wave packet version has the additional advantage of providing a more complete physical
description of the collision process, including causality. a
Our earlier steps of solving the scalar Helmholtz equation subject to the scattering boundary
condition | ψ (+) >→| ~k > +f eikr /r, were not in vain. The solutions to the stationary state
(+)
scattering problem ψ~ (along with any bound states) form a complete orthonormal (CON) set
k
that we shall use to construct the full wave packet description. Before taking that step, let us first
analyze the scattering process and the various dimensions involved in the real case.

7.1.1 Characteristic Lengths


To understand the scattering process in detail, let us return to the set-up involving a collimated
beam, a target, and detectors b . Again we deal with a dilute beam so that the scattering of a
single beam particle can be considered. For a thin target only one target particle is involved. (We
should really consider both the beam and target particles in appropriate wave packets; however, for
simplicity we shall consider wave packets for the incident particle only. The only issue here is to
deal with the relative motion wave packet.) Later we shall describe an ensemble of particles with
spin by a density matrix.
Now in the set-up several conditions are required in order for us to make a meaningful measure-
ment.
Detector
~k 0 :

  B

~k0 
- 
Beam -
-

Target

Figure 7.1: A collimated beam of particles strikes a target, scatters and then hits a detector. The
total distance travelled by this beam is denoted by L. Some of the collimated beam moves in the
forward direction (dashed lines), some scatters into the detector.
a Recent experimental developments have renewed interest in studying wave packets.
b This discussion is based on the Brandeis lectures of F. Low and on Rodberg & Thaler.
CHAPTER 7. SCATTERING-CONTINUED 147

First the beam must be collimated so that it strikes the target and reaches the detector only if the
target is present. When the target is removed the detector placed at θ > 0◦ should read zero. If that
were not true, how could we be sure anything we detect is really scattered rather than just received
directly from the accelerator? We can not therefore have an incident plane wave, which is spread
over all space and will trigger a detector even in the absence of a target. In addition to the slit, which
localizes the incident beam in the x-direction, the particle is localized in the incident z-direction in
the process of producing the beam (see later). Therefore, we consider for simplicity an incident
spherical wave packet c of radius ∆r as shown below with ∆r  L. Although this requirement that
the wave packet remains localized within the apparatus of length L, there is a limit to how small
∆r can be and still permit the setup to be a valid experiment; namely, one can not allow ∆r to be
so small that the packet spreads almost immediately from the incident slit region to the detector
region, thereby creating counts even without a target in place. We later impose an explicit condition
on ∆r to assure that the wave packet does not spread much over the apparatus dimension L.

.........................................
...........
.......................................
.........
...
.......... .............
Detector
........
....... ..........
...... Ψ
..... ......
sc B
.... ....
.... ....
*
 ... ...
 ~k 0
 ... ..
... ...
... ...
... ...
... ...
~k0 . .

Beam j - j - Φ
∆r Target ~k0

Figure 7.2: A particle in a spherical wave packet strikes a target, scatters and then hits a detector.
The total distance travelled by this beam is L. A forward direction wave packet Φ plus a spherical
shell wave packet Ψsc leave the target after/during the collision with the target. The target’s range
of interaction is denoted by b.

Another important requirement, which will allow us to study the energy dependence of scattering
processes, is that the beam have a precisely determined energy, at least as precise as is consistent
with the uncertainty principle and the collimation (∆x∆px ≥ h̄/2). Therefore the spread in momenta
(and hence in energy) of the incident particle must be small compared to the average (or group)
momentum k0 (~k0 = k0 ẑ)
∆k << k0
Recall that a free particle wave packet is formed by taking a superposition
Z i(~
k·r−ω(k) t)
e h̄ 2
Ψ(~r, t) = φ(~k − ~k0 ) d~k; ω(k) = k . (7.1)
(2π)3/2 2m

where φ(~k − ~k0 ) peaks at ~k0 with a width ∆k. The corresponding wave packet has a width ∆r ∼
(∆k)−1 and the above condition requires that

∆r >> k0−1 or ∆r >> λ


–0 .

We denote λ –0 = k0−1 as the deBroglie wave length ÷ by 2π (λ – = λ/2π). Hence the wave packet
has many oscillations inside the wave packet envelope as shown in the following figure. This re-
striction (on the lengths ∆r and λ –0 ) follows from our need to have a monochromatic (or nearly
monochromatic) beam, which we require to study the energy dependence of cross sections.
A third restriction needed to make the experimental set-up meaningful is that we can initially
isolate the incident particle from the target scattering region and indeed describe it as a free wave
packet (as given above). That requirement (and the need to separate out the detector) is partly
achieved by the requirement
∆r  L,
c The actual wave packet could be of different shape, for example of cylindrical or spheroidal shape, etc.
CHAPTER 7. SCATTERING-CONTINUED 148

<----------a------------>

Figure 7.3: A wave packet with envelope A, size ∆r ≈ a, and many internal oscillations of wave
–0 = k0−1 . Note that ∆r  λ
length λ –0 .

but we also need the scattering region to be small compared to L so that we are sure the process is
over before we detect it and not initiated until we have fully prepared the beam. Hence, we require
the scattering range, b, be small compared to macroscopic lab distances L

b  L.

Aside from the Coulomb force, the target scattering range is ∼ 10−8 cm for atoms (with shielded
Coulomb interactions) and ∼ 10−13 cm for nuclei, so the above condition is easily satisfied since L
∼ 1 meter say. We return to ∆r  L in a moment.
To be sure that we are really detecting an interaction between the target and beam particles, we
must also require that there be negligible spreading of the wave packet during the passage over a
distance L. If the incident particle had a rapid spreading it would reach the detector even without a
target; this condition is therefore simply a part of our original need for a collimated beam over the
macroscopic distance L. Recall that our analysis of a free particle wave packet led us to
~
Ψ(~r, t) = ei(k0 ·~r−ω0 t) A(~r − ~v0 t) + · · ·
h̄ 2
with ω0 = 2m k0 and a group velocity ~vg = ~v0 = ∇~ k ω(k) |k . The higher order terms (the spreading
0
terms) were negligible only if
t h̄2 1
= 1
h̄ 2m (∆r)2
L
or since t ∼ v0 we had obtained the following condition for negligible spreading

1 L λ–0
=  1.
4 ∆r ∆r
Since the deBroglie wave length is already assumed much smaller than the wave packet width
–0  ∆r), the above condition limits how small ∆r can be, i.e. the ratio of L to ∆r. Some typical

values for these quantities are:
L ' 1 meter = 102 cm
∆r ' 10−1 cm
k0−1 = λ
–0 ' 10−8 cm
∆r 7
–0 ' 10
λ
1 3
' 10 × 10−7 ' 10−4 << 1!
4
CHAPTER 7. SCATTERING-CONTINUED 149

∆r
' 10−3 ∆r < L
L
b ' 10−8
b
' 10−10 b<L
L
Thus all conditions are satisfied and an actual measurement can be made. Here we deal with an
important aspect of Quantum Mechanics; namely, that only certain setups qualify as an experiment,
which is the practical realization of Bohr’s complementarity principle.
As a final remark about the characteristic lengths, let me point out that the incident wave
packet size ∆r typically has a length and width of different magnitudes with for example w ≤ `,
with ` ∼ 10−3 cm.

~k0
Beam w -
` Target

Figure 7.4: A particle in a rectangular wave packet of length ` and width w is prepared.

In part, the packet lengths w and ` are fixed by the preparation, slit and injection steps. The
width is determined partly by the collimator, a typical value for w is w ∼ 10−4 cm. The length is
determined by the accelerator; one must inject particles from an ion source which gives an initial
wave packet whose size is determined by the collision times and by the radio frequencies used. d A
typical value of ` is ∼ 10−3 cm. For simplicity, we assume ` ∼ ω ; ∆r, and consider spherical wave
packets. We will see later that the definition of the cross section is independent of the detailed shape
of the wave packet, provided we have a valid experimental setup, e.g. the characteristic lengths are
all properly designed.
One point worth noting concerns the condition ∆k << k0 , which is claimed here to arise from
our need for a monochromatic beam. In addition to the variation in energy for each packet, we
should really consider the variation in k0 (the group momentum) for an ensemble of packets. It is
really ∆k0 over such an ensemble of packets that determines the accelerator’s characteristic energy
fluctuation. This is an accelerator design feature.

7.2 Scattering Wave Packets


(Ref: Rodberg and Thaler - Chapter 2-on reserve)
Now we can construct appropriate wave packets that incorporate all of the above restrictions on
the characteristic lengths. The basic step is to replace the complete orthonormal set of plane wave
(+)
stationary states by the complete orthonormal set ψ~ (plus bound states)
k

~
eik·~r (+)
→ ψ~ (r).
(2π)3/2 k

Since we will be concerned about large distances (r  b = size of bound state), the bound state
part can be dropped for formation of the asymptotic wave packet.
The property of being a complete orthonormal set is expressed by
Z
(+)∗ (+) (+) (+)
d~r ψ~ (~r ) ψ~ 0 (~r ) = δ(~k − ~k 0 ) ≡< ψ~ | ψ~ 0 >,
k k k k

d After acceleration the length is decreased somewhat by relativistic contraction. For high energies (and heavy ions)

the wave packet takes on a pancake shape.


CHAPTER 7. SCATTERING-CONTINUED 150
Z
∗ (+)
ψB (~r ) ψ~ (~r ) d3 r = 0, (7.2)
k

and in closure form as


Z
(+) (+)∗
X
ψBn (~r )ψBn (~r 0 ) + d~k ψ~ (~r ) ψ~ (~r 0 ) = δ(~r − ~r 0 ).
k k
n
(−)
Note that ψ~ plus the bound states form another, independent complete orthonormal set.
k
Using the above properties, we can build our wave packet using the complete orthonormal set
ψ (+) in place of the plane wave set.
Z
(+) i
Ψ(+) (~r, t) = φ(~k − ~k0 ) ψ~ (~r )e−iω(k)t d~k = e− h̄ Ht Ψ(+) (~r, 0). (7.3)
k

Again, bound state terms are omitted here since they vanish for large r. Note Ψ(+) (~r, t) satisfies
∂ (+) (+)
(H − i ∂t )Ψ(+) = 0, since Hψk = h̄ω(k) ψk .
(+)
When the packet is far from the target region, we can replace ψ~ by its asymptotic value; this
k
follows from the fact that ∆r < L and we can separate out the incident beam for earlier times. Thus
for large r,
φ(~k − ~k0 ) i~k·~r eikr −iω(k)t
Z
Ψ(+) (~r, t) → d~k 3/2
[e + fk (θ, φ) ]e
(2π) r
= Φ(~r, t) + Ψsc (~r, t)
Z ~ ~
φ(k − k0 ) i~k·~r −iω(k)t ~
Φ(~r, t) = d~k e e = ei(k0 ·~r−ω0 t) A(~r − ~v0 t) + · · · (7.4)
(2π)3/2
~
with ~v0 = ∇ω(k) |k0 = h̄k
m ẑ = v0 ẑ; Φ is the plane wave packet, which propagates as before. (See
0

Lectures last term; now we call A(r, t) |t=0 ≡ A(~r ).)


The scattered wave packet is
φ(~k − ~k0 ) eikr −iω(k)t
Z
Ψsc (~r, t) = d~k f k (θ, φ) e . (7.5)
(2π)3/2 r
Now we will show that for t < 0 there is no scattered wave. Here we set t = 0 at the time of full
impact between the incident wave packet and the target, i.e. where the maximum of A overlaps
with the target particle.
For earlier times t = −(L/2)/v0 , the incident packet is far from the target region and we indeed
expect Ψsc → 0 as t → −(L/2)/v0 .
Proof

We assume fk (θφ) = fk0 (θφ) + (k − k0 )( ∂k fk (θφ) |k0 ) + · · · and assume that over the width ∆k of
~ ~
φ(k − k0 ), the amplitude fk varies slowly. That is: the energy dependence of fk is slow compared to
the relatively sharp peak in φ(~k − ~k0 ). For example ∆E(wavepacket) ' 10−5 ∆E(resonance width).

Energy

Figure 7.5: As a function of energy the envelope function φ is seen to be narrow compared to the
relatively slow variation of the scattering amplitude fk (θ, φ) with energy.

After dropping the second term in the expansion of the amplitude e , we have
φ(~k − ~k0 ) ikr −iω(k)t
Z
fk (θ, φ)
Ψsc (~r , t) = 0 d~k e e . (7.6)
r (2π)3/2
e Later we will include that term and obtain the Wigner causality condition on the energy dependence of the

scattering amplitude
CHAPTER 7. SCATTERING-CONTINUED 151

Now we use:
~k = ~k0 + ~k − ~k0 ,

and
| ~k |2 = k 2 = k02 + 2~k0 · (~k − ~k0 ) + (~k − ~k0 )2 = −k02 + 2~k0 · ~k + (~k − ~k0 )2
or
~k − ~k0 (~k − ~k0 )2 1/2
k = k0 [1 + 2k̂0 · + ]
k0 k02
2k̂0 · (~k − ~k0 ) 1/2
' k0 [1 + ] ' k0 + k̂0 · (~k − ~k0 ),
k0
where inside the integral we have used the condition ∆k ≤| ~k − ~k0 |<< k0 required to have a
monochromatic beam. Thus

φ(~k − ~k0 ) i(~k−~k0 )·k̂0 r −iω(k)t


Z
fk0 (θφ) ik0 r
Ψsc (~r, t) = e d~k e e .
r (2π)3/2

Now for ω(k) we use

h̄ 2 h̄
ω(k)t = k t= (−k02 + 2~k0 · ~k + (~k − ~k0 )2 ) t
2m 2m
h̄ 2 h̄ h̄
'− k t + ~k0 · ~k t + (∆k)2 t
2m 0 m 2m
' −ω0 t + ~v0 · ~k t +  ' −ω0 t + ~v0 · ~k t ' +ω0 t + ~v0 · (~k − ~k0 ) t.
The small spreading condition is again
h̄ 1 L λ–0
(∆k)2 t== << 1,
2m 4 ∆r ∆r
for | t |' L/v0 . We can express the scattered wave:

eik0 r −iω0 t φ(~k ) i~k·(k̂0 r−~v0


Z
Ψs (~r, t) = fk0 (θ, φ) e d~k e t)
r (2π)3/2

eik0 r −iω0 t
= fk0 (θ, φ) e A(k̂0 r − ~v0 t). (7.7)
r
Note: For φ(~k − ~k0 ) → δ(~k − ~k0 ), this expression returns us to the stationary state case with
(+)
Ψ(+) (~r, t) = ψ~ (~r)e−iω(k0 )t
k0

and
1
A(k̂0 r − v0 t) =
(2π)3/2
and
eik0 r e−iω0 t
Ψsc (~r, t) = fk0 (θ, φ) .
r (2π)3/2
So all is in order.
Returning to the case of a general peaked function φ, we have found that for large ~r

~ fk0 (θ, φ)
Ψ(+) (~r, t) = e−iω0 t {A(~r − ~v0 t) eik0 ·~r + A(k̂0 r − ~v0 t)} (7.8)
r
Note: a very important fact: a vector ~r appears in the first term, but in the second term we have
| ~r | a magnitude in the envelope functions. Also since k̂0 = ẑ and ~v0 = v0 ẑ, the amplitude of
the scattered wave is simply given by A(ẑ(r − v0 t)). Consider for a moment a typical wave packet
CHAPTER 7. SCATTERING-CONTINUED 152

r2
envelope–a Gaussian: A(~r) = (πw21)3/4 e− 2w2 , with w denoting the 1/2-width. The main point is a
spherical wave packet depends on | ~r | or r2 only. Therefore, for such a Gaussian packet

A(ẑ(r − v0 t)) = A(r − v0 t)

with r a magnitude and hence always positive. Now for earlier times, way before the collision we
have t = −L/2/v0 , and hence v0 t is negative and the envelope function is A(r+ | v0 t |) = A(r + L2 ).
Since r + L/2 is large and positive the amplitude of the scattered wave is (essentially) zero. We
have therefore shown that for times before the overlap of the beam and target particle wave packets
( t  0), no scattered wave occurs. An outgoing scattered wave starts to develop only when the
front edge of the wave packet enters into the potential range b. That makes good physical sense! It’s
causality.
We have for t  0
~
Ψ(+) (~r, t) → Φ(~r, t) = e−iω0 t eik0 ·~r A(~r − ~v0 t).
For large r and for positive times t > 0; t → L/2v0 , we have enough time for the collision to be
completed and the limit is

fk0 (θ, φ)
Ψ(+) (~r, t) → Φ(~r, t) + ei(k0 r−ω0 t) A(r − v0 t).
r
After the collision (t → L/2v0 or ∞) the waves look like a free wave packet Φ plus an outgoing
spherical wave packet shell. The spherical shell represents an outgoing wave packet of thickness ∆r,
with its angle dependence given simply by fk0 (θ, φ) and with the usual 1/r fall-off. The wave packet
Φ is represented by the open circle region; it is moving to the right with velocity ẑv0 .

.........................................
...........
.........
...................................... ...
.......... .............
........
....... ..........
......
Ψ
..... ......
.... .... sc B
.... ....
*

 .... ..
 ... ....
... ...
... ...
... ..
... ...
~k0 .

Beam j - j j - Φ
Target ~k0

Figure 7.6: For earlier times t < −b/v0 there is just a incident spherical wave packet ball, at t = 0
the spherical wave packet is centered at the potential of range b. The collision is finished when
t > b/v0 and thereafter we have the spherical wave packet Φ moving along the ẑ axis plus an
outgoing spherical shell wave packet Ψsc of thickness ∆r.

The important point here is that Φ is confined to a small region along the z–axis–as determined
by the wave packet size ∆r. Hence, aside from the forward direction, the outgoing spherical wave and
the incident beam are separated by the non-overlap of their wave packets. In the forward direction,
they do overlap and interfere to deplete the forward direction flux. That brings us back to the
optical theorem, which we will re-derive in a moment.
First let me use the clean separation of the spherical or scattered waves for θ > 0, to define the
cross-section in a more rigorous fashion.
CHAPTER 7. SCATTERING-CONTINUED 153

7.2.1 Cross Section & Wave Packets


Since we can neglect the spreading current (see Lecture last term), and the θ > 0 scattered wave is
separated by the non-overlap of its wave packet with Φ, we can evaluate the probability of receiving
a particle at the detector placed at a fixed distance r from the target and at the angles θ, φ of size
dA by Z ∞
dPD = | Ψsc (~r, t) |2 v0 dt dA
−∞
Z ∞
2
= v0 dΩ | fk0 (θ, φ) | dt | A(r − v0 t) |2 . (7.9)
−∞

This quantity depends on the wave packet’s shape (i.e. on A), but we define the cross section using
the incident flux density integrated over all time
Z ∞ Z ∞
h̄k0 particles
I~ = dt ~jincident = ẑ dt | A(ẑr − ~v0 t) |2 = .
−∞ m −∞ cm2

We evaluate A at ~r = ẑr corresponding to the forward direction. Hence using A(ẑr−~v0 t) = A(r−v0 t)
and v0 = h̄k0 /m the common wave packet integral cancels out

dPD
dσ ≡ = dΩ | fk0 (θ, φ) |2 ,
| I~ |

where the dependence on the wave packets details has dropped out. Thus we return to our old result
obtained using the stationary state approach

=| fk0 (θ, φ) |2 in cm2 , (7.10)
dΩ
where fk0 is obtained solving the stationary state scattering problem. The above result of course is
valid only for the several assumption used concerning the characteristic lengths, non-spreading, and
slow variation of fk compared to φ(~k − ~k0 ).

7.3 Optical Theorem & Wave Packets


The clean separation provided by the wave packets permits us to isolate the scattered wave from
the incident beam at all angles except for the forward direction. The interference between Ψsc and
Φ accounts for the flux lost from the forward direction by scattering (and reactions). The overall
conservation of flux will now again lead us to the optical or shadow theorem, except we will be
more rigorous. The wave packet fall–off in r provides the justification for dropping the O(1/r) and
O(1/r2 ) terms in our earlier stationary state rendition.
The proof is as follows:
ih̄Ψ̇ = HΨ
~ · ~j = −ρ̇, provided we have a hermitian potential V † = V, and hence no source or sink
leads us to ∇
f
of flux.
Integrating over a surface at infinity which includes all of the system
Z Z
∂ ~=0
d~r ρ = − ~j · dA
∂t
no net flux lost. WeRassume the time is limited to values for which all of the system is within that
large surface. Thus d~r ρ(~r, t) = constant. Evaluate that integral for t → ±∞ and equate as
Z Z Z
d~rρ(r, −∞) = d~r | Φ(~r, t) |2−∞ = d~r | Φ(~r, t) + Ψsc (~r, t) |2+∞

f We could generalize the following prove to include such creation and annihilation effects by considering the

equation with a source term ∇ ~ · ~j + ρ̇ = S(~


r, t), and by extending the definition of the total cross section to be the sum
of all elastic, inelastic and absorption and reaction events. For simplicity we shall consider the case of S(~ r, t) = 0.
CHAPTER 7. SCATTERING-CONTINUED 154
Z Z Z
= d~r ρ(~r, +∞) = d~r | Φ(~r, t) + d~r [ Φ∗ (~r, t) ψsc (~r, t) + Φ(~r, t) Ψ∗sc (~r, t) |+∞
|2+∞
Z
+ d~r | Ψsc (~r, t) |2+∞ .

An important aspect of this proof is that we include use the full forms for both Φ and Ψsc .
Thus the spreading of both the incident and the scattered wave packets are incorporated, which is
essential for the full proof presented here (see Rodberg & Thaler).
~
Note that using Φ(~r, t) = e−iω0 t eik0 ·~r A(~r − ~v0 t), we have:
Z Z
d~r | Φ(r, −∞) |2 = d~r | Φ(r, +∞ |2 ,

and therefore
Z Z
d~r | Ψsc (~r, t) |2+∞ = − d~r [ Φ∗ (~r, t) ψsc (~r, t) + Φ(~r, t) Ψ∗sc (~r, t) |+∞ .

and
Z Z
d~r | Ψsc (~r, t) |2t→+∞ = −2Re d~r Φ(~r, t) Ψ∗sc (~r, t) |t→+∞ .

The equation above is a statement that destructive interference between the final state spherical
wave packet moving along the z-axis and the part of the spherical shell wave packet also along the
z-axis account for the total cross section. Since the interference is destructive, flux lost from the
θ = 0◦ region must simply reappear elsewhere; namely, in θ > 0◦ regions. The loss of flux at θ = 0◦
corresponds to the formation of a shadow, and hence the optical theorem is often called the shadow
theorem. We know from classical optics that the energy missing from the shadow appears in other
illuminated regions, so this result is a quite familiar experience.
The pure scattering term is:
Z Z Z ∞ Z ∞
d~r | Ψsc (~r, t) |2t→+∞ = | fk0 (θ, φ) |2 dΩr̂ dr | A(r − v0 t) |2 = σT OT dr | A(r − v0 t) |2
0 0

and we have
Z ∞ Z
σT OT 2
dr | A(r − v0 t) | = −2 Re d~r [ Φ(~r, t) Ψ∗sc (~r, t) |2t→+∞ .
0

Since the Ψsc is essentially independent of angle near the forward direction, the right hand side
(RHS) of this equation becomes:
Z Z
2 ∗
RHS = −2 Re drr Ψsc (r, +∞) dΩr̂ Φ(~r, +∞) .

Now the spherical wave packet A(~r − ~v0 t) and the spherical shell wave packet A(r − v0 t) overlap
only in the forward direction. As t → +∞ the solid angle integration of d~r is extremely small, so
we can evaluate fk0 (θ, φ) at θ = 0. Thus
Z ∞ Z ∞ Z
2 ∗ −i(k0 r−ω0 t) ∗
σT OT dr | A(r − v0 t) | = −2 Re{fk0 (θ = 0) r dr e A (r − v0 t) dΩr̂ Φ(~r, +∞) .
0 0

The only part of the right hand side (RHS) which depends on angle involves

d3 k
Z Z Z
dΩr̂ Φ(~r, +∞) = φ(~k − ~k0 ) dΩr̂ ei(~k·~r−ωk t) .
(2π)3/2
Since Z
~ sin(kr)
dΩr̂ eik·~r = 4π ,
kr
we have Z Z
1 4π −iωk t
dΩr̂ Φ(~r, +∞) = d3 k φ(~k − ~k0 ) e sin(kr),
(2π)3/2 kr
CHAPTER 7. SCATTERING-CONTINUED 155

and
d3 k
Z Z
4π −iωk t +ikr
dΩr̂ Φ(~r, +∞) = φ(~k − ~k0 ) e (e − e−ikr ) (7.11)
(2π)3/2 2ikr
4π i(k0 r−ω0 t)
≈ [e A(r − v0 t) − e−i(k0 r+ω0 t) A(r + v0 t)] |t→+∞
2ik0 r
4π i(k0 r−ω0 t)
≈ + e A(r − v0 t) |t→+∞
2ik0 r
where we used A(r + v0 t) |t→+∞ = 0. Inserting this result into our prior equation, we obtain
Z ∞ Z ∞
2π ∗
σT OT dr | A(r − v0 t) |2 = −2Re{fk∗0 (θ = 0) dr A (r − v0 t)}A(r − v0 t)},
0 0 ik0
g
from which we deduce the optical theorem

σT OT = Im[fk0 (θ = 0)].
k0
Note that since the total cross section is greater than zero, this relation requires that the forward
direction elastic scattering amplitude fk0 (θ = 0) must have a positive imaginary part. In the general
version of this theorem the total cross section is a sum of the elastic, inelastic, absorption, reaction
· · · contributions and the RHS has the forward scattering elastic amplitude. It can therefore be used
to determine if all of the possible physical processes have been detected or that something is missing.

g Peierls has an interesting rendition of the original derivation of this result, see page 6 of “Surprises in Theoretical

Physics.”.
CHAPTER 7. SCATTERING-CONTINUED 156

7.4 Partial Wave Expansion


Two-body scattering involves a six-dimensional problem for the wave function Ψ(~r1 , ~r2 , t). One
should really deal with wave packets for both the target and beam particles. However, since the
center of mass motion separates out, the wave packet splits into a c.m. packet times a relative
motion packet. It is the latter relative motion packet that was considered in the previous section
(with the reduced mass to be included).
The mathematics involved in the c.m. extraction is:
~ t) Ψrel (~r, t).
H = Hc.m. + Hrelative and Ψ = χc.m. (R,

where χc.m. and Ψrel are the wave packet of the c.m. and relative motion, respectively. We have
separate Schrödinger equations

ih̄χ̇cm = Hcm χcm and ih̄Ψ̇rel = Hrel Ψrel .

Thus the 6-dimensional problem is reduced to two 3-dimensional wave packet problems. The
relative motion problem for scattering involves solving the scalar Helmholtz equation subject to the
stationary state boundary condition

~ eikr
ψ (+) (r) → (2π)−3/2 {eik·~r + f (~k 0 , ~k) }.
r
There results
4π 2 µ ~ 0 (+)
f (~k 0 , ~k) = − < k | V | ψ~ >,
h̄2 k

with reduced mass µ = mM 1 m2


.
The center of mass reduction has helped, but we still have a three dimensional problem to
solve. Hence, we need to introduce the partial wave expansion to reduce the problem to three,
one-dimensional problems. As for the case of bound states, we find for central potentials the in-
troduction of eigenstates of orbital angular momenta Y`m (θφ) reduces the scattering problem to a
one-dimensional radial equations for each value of `. The number of important `-values is deter-
mined by the energy and size of the scattering potential. For orbital angular momenta ` ≤ kb (b is
the range of V), there will be contributions from the potential. For large collision orbital angular
momenta ` > kb, the collision will be outside of the potential range (due to the centrifugal barrier
effect) and the radial problem will be essentially unchanged by V(r) and remain the free particle
solution. These assertions will be shown later. For now it is important to know that the number
of radial equations is, for lower energies and small target sizes, sufficiently small that one gains by
introducing a partial wave expansion. For high energies and/or long range potentials (Coulomb
potential, for example) the use of a partial wave expansion must be reconsidered since it will not
necessarily be advantageous. (High energy approximations, such as the eikonal method, avoid the
need to use a partial wave series.)

7.4.1 Transformation from Center of Mass to Laboratory Systems


There is one aspect of the use of relative motion to describe scattering that requires some discussion
before proceeding to the partial wave series. The obvious point is that experiments are performed in
the laboratory system, which usually consists of having the target at rest. (The exception is when
we have clashing or colliding beams for which the laboratory and c.m. systems are identical for equal
massed beams). The scattering amplitude is obtained (and the phase shifts as defined later) in the
relative or center of mass frame. It is therefore often necessary to transform our expressions for
the differential cross sections from the center of mass (c.m.) frame to the corresponding laboratory
frame (lab). We have:
CHAPTER 7. SCATTERING-CONTINUED 157
0 0
 p~1 = m1~v1L

- s - P~ = m1~v1 = p~1 = M V
~cm
p~1 p~2 @
@
R p~20 = m2~v2L
@ 0

Figure 7.7: The laboratory system in which the target particle has zero momentum p~2 = 0 as
indicated by the dot. The total momentum is P~ and the velocity of the center of mass is V ~cm . The
0 0
final velocity in the laboratory system are ~v1L and ~v2L for particles 1 & 2, respectively
0
p~12


- s 
p~12 −~
p12
0
−~
p12

Figure 7.8: The center of mass, or relative motion, system in which the total momentum P~ = 0.
0
The initial momenta in this system are p~12 and −~
p12 . The final momenta in this system are p~12 and
0
−~
p12 .

m2 p~1 − m1 p~2
p~12 = = µ(~v1 − ~v2 ) = m1~v1c = µ~viL
M

~ = m1~v1L
V
M
To relate these vectors, we use the fact that the center of mass motion is seen by moving with V
in the laboratory; i.e. we add vector velocities as ~v1c + V ~ = ~v1L and ~v1c + V
~ = ~v2L . Note that: ~v1c
is the velocity of particle 1 in the cm system, ~v1L is the velocity of particle 1 in the laboratory(L)
system, (~v1 − ~v2 ) is the relative velocity of particles 1 and 2, and V~ is the velocity of the center
of mass as seen in the laboratory system. The addition rule is describes by the following figure,
which allows us to define a laboratory θL and a cm θcm scattering angles. The azimuthal angles are
unchanged φL ≡ φcm = φ. We get:

*



0 0
~v1L   ~v1c


  θL - θc
~
V
Figure 7.9: Addition of the center of mass velocity to the final relative velocity of particle 1 to obtain
the final laboratory momentum of particle 1.

0 0
vc1 cos θc + V = vL1 cos θL (7.12)
φL = φc
0 0
vc1 sin θc = vL1 sin θL
CHAPTER 7. SCATTERING-CONTINUED 158

sin θc
tan θL =
cos θc + γ
m1
V M |~v1L | m1
γ ≡ 0 = m2 =
vc1 M | ~
v 1L | m2

(for elastic scattering) Note we have used µ ≡ mM1 m2


and m1~v1c = µ~v1L . Corresponding relativistic
transformations, using four-vectors, reduce to these nonrelativistic results.

Cross Section Transformation


The differential cross section thus transforms as
dσ dσ
( )cm dΩcm = ( )lab dΩlab ,
dΩ dΩ
which states that the same number of counts are scattered to dΩL as to dΩcm as determined by the
detector. Therefore,
dσ dΩcm dσ dΩcm
( )lab = σL (θL ) = ( )cm = σcm (θcm )
dΩ dΩlab dΩ dΩlab
or
d cos θcm dµc
σL (θL ) = σcm (θcm ) = σcm (θcm ).
d cos θlab dµL
A little algebra is required to get
dµc d cos θC
=
dµL d cos θL
First note that with µL ≡ cos θl L, and µc ≡ cos θc ,

(1 − µ2L )1/2 tan θc (1 − µ2c )1/2


tan θL = = =
µL cos θc + γ µc + γ
or
1 (1 − µ2c ) (µc + γ)2 + (1 − µ2c )
2 =1+ 2
=
µL (µc + γ) (µc + γ)2
| µc + γ |
µL =
(2µc γ + γ 2 + 1)1/2
which we can use to generate the requisite derivative

dµL 1 1 | µc + γ |
= − (2γ)
dµc (2µc γ + γ 2 + 1)1/2 2 (2µc γ + γ 2 + 1)3/2
(2µc γ + γ 2 + 1)− | µc + γ | γ
=
(2µc γ + γ 2 + 1)3/2
=| (µc γ + 1) | /(2µc γ + γ 2 + 1)3/2
Therefore
(1 + γ 2 + 2γ cos θcm )3/2
σlab (θlab ) = σcm (θcm )
| 1 + γ cos θcm |
relates differential cross section in the c.m. to the differential cross section measured in the lab
system. The total cross sections are the same in both frames since the net counts observed at all
angles is the same!
... σ sc (E)
same in lab and c.m.
This is a non-relativistic treatment. Corresponding relations can be derived for relativistic colli-
sions.
CHAPTER 7. SCATTERING-CONTINUED 159

Examples of transformations.
One special case is for equal masses, say two nucleons or two electrons. Then

γ=1

sin θc 2 sin θ2c cos θ2c θc


tan θL = = θ
= tan
cos θc + 1 2 cos 22 c 2
Therefore, for equal masses
θc
θL =
2
and
σlab (θL ) = 23/2 (1 + cos θc )1/2 σc (θc ) = 4 cos θL σc (θc )
For pure S−waves σc is independent of θc and

σlab (θL ) ∼ cos θL

and θL ≤ π/2 resulting in forward scattering in the lab for isotropic elastic scattering in the cm.

7.4.2 Plane Wave Decomposition


The radial equation for a central potential yields solutions ψ = r−1 u` (r)Y`m (θφ), which in general
provide a complete orthonormal set of radial (u` /r) and angular functions (Y`m ) (provided all bound
state and continuum states are included). For the special case of V = 0, there are only continuum
solutions, which were seen earlier to be spherical Bessel functions. Thus j` (kr)Y`m (θφ) provides a
complete orthonormal set of functions, which are finite everywhere (this finite condition being the
reason no irregular spherical Bessel functions n` (kr) are needed for the CON set). It is therefore
not surprising that the free particle solution of a plane wave can be decomposed as
~ r
eik·~r 2X ` ∗
= i j` (kr)Y`m (r̂)Y`m (k̂), (7.13)
(2π)3/2 π
`m

or
~
X
eik·~r = i` (2` + 1)j` (kr)P` (cos θkr ),
where the relation
4π X ∗
P` (cos θkr ) = Y`m (r̂)Y`m (k̂),
2` + 1 m
was used.
From the asymptotic properties of the spherical Bessel function:
sin(kr − `π/2)
j` (kr) −→ (7.14)
kr→∞ kr
cos(kr − `π/2)
n` (kr) −→ −
kr→∞ kr
ikr−`π/2
e
h` (kr) −→ ,
kr→∞ ikr
(7.15)

with the spherical Hankel function defined by: h` (kr) = j` (kr) + in` (kr), one finds the plane wave,
for large distances r, is composed of incoming and outgoing spherical waves of equal amplitude:
~ r
eik·~r 2 X i` ei(ρ−`π/2) e−i(ρ−`π/2) ∗
→ [ − ]Y`m (r̂)Y`m (k̂). (7.16)
(2π)3/2 π 2i ρ ρ

Note these incoming and outgoing radial waves do not arise from the scattering process but are
already in the free plane wave motion - they arise from the partial wave decomposition. This build-
up of a plane wave as given above is the most important step in the partial wave decomposition.
CHAPTER 7. SCATTERING-CONTINUED 160

7.4.3 Stationary State Expanded in Partial Waves


The plane waves (or free particle) wave function has been seen to be decomposed into radial terms
which are, for large r, both incoming and outgoing spherical waves. The solution of the stationary
state scattering problem includes this plane wave plus the effects of the potential, i.e. of the scat-
tering. We will see later that the presence of a potential serves to modify the expansion given by
the plane way; in the simplest case of elastic scattering one finds a change in phase of the outgoing
radial wave terms of the partial wave expansion. This modification, and the presence of the plane
wave, suggests that the stationary state solutions be expanded in a form similar to that of a plane
wave: r
(+) 2 X ` (+) ∗
ψ~ (~r) = i ψ` (kr)Y`m (r̂) Y`m (k̂). (7.17)
k π
`m
(+)
Now the radial functions ψ` (kr)
[plus the corresponding bound state solutions] form a complete
orthonormal set of radial functions. For a plane wave only, ψ` reduces to
~
(+) eik·~r
ψ` −→ j` (kr) gives ψ (+) (~r ) −→ .
(2π)3/2

Green’s Function Expansion


Now we can use the plane wave expansion to obtain a decomposition of the full out- going wave
Green’s function given by
~0 0
eik ·(~r−~r ) ~ 0
Z
1
(~r | G(k 2 + i) | ~r 0 ) = dk (7.18)
(2π)3 k 2 + i − k 02
1 eikR
= −
4π R
2 X ∞ dk 0 k 02 j` (k 0 r) j` (k 0 r 0 )
Z

= 2 + i − k 02
Y`m (r̂) Y`m (r̂0 )
π k
`m 0
X
= (r | G` (k 2 + i) | r0 ) Y`m (r̂) Y`m

(r̂0 )
`m
X (2` + 1)
= (r | G` (k 2 + i) | r0 ) P` (cos θrr0 ).

`

We have defined the Green’s function in the `th partial wave by

2 ∞ dk 0 k 02 j` (k 0 r)j` (k 0 r0 )
Z
2 0
(r | G` (k + i) | r ) = −
π 0 k 02 − k 2 − i

= −ki j` (kr<) h` (kr>),


where r< > r> .

Proof of last equation


Consider the form

dk 0 k 02 j` (k 0 r) j` (k 0 r0 )
Z
2 0 2
(r | G` (k + i) | r ) = − . (7.19)
π 0 k 0 − k 2 − i

We can extend the integration over k to the full −∞−+∞ region since each spherical Bessel function
has the property j` (−kr) = (−1)` j` (kr) and (−1)2` = +1

1 ∞ dk 0 k 02 j` (k 0 r) j` (k 0 r0 )
Z
(r | G` (k 2 + i) | r0 ) = − . (7.20)
π −∞ k 0 − k 2 − i
CHAPTER 7. SCATTERING-CONTINUED 161

Now take r > r0 which is denoted by r ≡ r> and r0 ≡ r<. Replacing j` (k 0 r0 ) → j` (k 0 r>) →
1 0 ∗ 0
2 [h` (k r>) + h` (k r>)]. Now introduce Hankel functions to prepare for a contour integration


dk 0 k 02
Z
1
= − 2 − i `
[h (k 0 r>) + h∗` (k 0 r>)]j` (k 0 r<). (7.21)

−∞ − kk 02
Z ∞ Z ∞
1 dk 0 k 02 h` (k 0 r>)j` (k 0 r<) 1 dk 0 k 02 h` (k 0 r>) j` (k 0 r<) ∗
= − 02
+ [− ]
2π −∞ k − k − i 2 2π −∞ k 02 − k 2 + i
= I1 + (I2 )∗ ,

where we have defined ∞


dk 0 k 02 h` (k 0 r>)j` (k 0 r<)
Z
1
I1 ≡ −
2π −∞ k 02 − k 2 − i
and ∞
dk 0 k 02 h` (k 0 r>) j` (k 0 r<)
Z
1
I2 ≡ −
2π −∞ k 02 − k 2 + i
The spherical Hankel function has the asymptotic form

ei(ρ−`π/2)
h` (ρ) −→ ,
ρ→∞ iρ
which indicates that both integrals above can be extended to contour integrations, since we get zero
0
contribution from the upper ∞ circle, i.e. eik (r>−r<) → 0 on infinite circle k 0 → i∞.
...........................................
.....
............ .. ... ..........
......... ........
.
......... .....
.. .....
....
....
...  ........
.. R ...
... ...
.. ...
.... ...
...
..
... ...
.... -r × ...
. .. . - .
−∞ ∞

Figure 7.10: The complex k 0 plane with the contour closed in the upper half complex plane, thereby
including the pole or singularity in the out-going partial wave Green’s function (r | G` (k 2 + i) | r0 )
at k 0 = k + i 2|k|

pole for I1 as marked by a ×, and the k 0 = −k + i 2|k|

pole for I2 is marked by a
dot. The infinite upper circular contour’s radius is denoted by R.

The first contour integral I1

dk 0 k 02 h` (k 0 r>)j` (k 0 r<) 2πi k 2


I
1 k
− 02 2
= − h` (kr>)j` (kr<) = −i j` (kr<) h` (kr>)
2π k − k − i 2π 2k 2

has upper half plane poles at k 0 = k + i


k, whereas, the second contour integral I2

dk 0 k 02 h` (k 0 r>) j` (k 0 r<) 2πi k 2


I
1 k
− 02 2
=− h` (−kr>)j` (−kr<) = +i j` (kr<) h∗` (kr>)
2π k − k + i 2π −2k 2
h
has upper half plane poles at k 0 = −k + i
k . These are the only poles, since the spherical Bessel
functions are analytic functions. We have used the residue theorem and now obtain

k k
(r | G` (k 2 + i) | r0 ) = −i h` (kr>)j` (kr<) + [+i h∗` (kr>)j` (kr<)]∗ = −ik j` (kr<) h` (kr>),
2 2
which completes the proof. Note that this result can also be obtained by the “ space-ordered”
Wronskian approach (see Homework.).
h Note that j` (−kr<) = (−1)` j` (kr<) and h` (−kr>) = (−1)` h∗` (kr>)
CHAPTER 7. SCATTERING-CONTINUED 162

Incoming and Standing Wave Boundary Condition


The incoming wave boundary condition Green’s functions is:

2 ∞ dk 0 k 02 j` (k 0 r)j` (k 0 r0 )
Z
(r | G` (k 2 − i) | r0 ) = − (7.22)
π 0 k 02 − k 2 + i
= +k i j` (kr<) h∗` (kr>)

Note G0 (E − i) = G†0 (E + i).


The standing wave boundary condition Green’s functions is:
Z ∞ 0 02
2 dk k j` (k 0 r)j` (k 0 r0 )
(r | G` (k 2 ) | r0 ) = − P (7.23)
π 0 k 02 − k 2
= +k j` (kr<) n` (kr>).
1
Note (G0 (E − ie) + G0 (E + ie))/2 = P E−H 0
.

7.4.4 Partial Wave Expansion


Phase Shifts
We have expanded the plane wave in a complete orthonormal set of radial and angle functions
j` (kr)Y`m (θφ) and introduced corresponding decompositions of the full wave functions and of the
Green’s function for outgoing waves
~ r
eik·~r 2 X i`
= [h` (kr) + h∗` (kr)]Y`m (r̂)Y`m

(k̂)
(2π)3/2 π 2
`m
r
2X ` ∗
= i j` (kr)Y`m (r̂)Y`m (k̂)
π
`m
r
(+) 2 X ` (+) ∗
ψ~ (~r) = i ψ` (kr)Y`m (r̂)Y`m (k̂). (7.24)
k π
`m

We have shown that the Green’s function is decomposed by


X
(~r | G(k 2 + i) | ~r0 ) = (r | G` (k 2 + i) | r0 )Y`m (r̂)Y`m

(r̂0 ) (7.25)

and
(r | G` (k 2 + i) | r0 ) = −ki j` (kr<)h` (kr>). (7.26)
Using these relations, it follows that the integral equation in the `th partial wave consists of an
incident wave term plus a scattered wave:
Z ∞
(+) (+)
ψ` (kr) = j` (kr) + (r | G` (k 2 + i) | r0 )r02 dr0 U (r0 ) ψ` (kr0 ). (7.27)
0

For large r, with a potential U (r) = 2m


h̄2
V (r) of finite range (r → r> and r0 → r<), the asymptotic
wave function is:
Z ∞
(+) (+)
ψ` (kr) −→ j` (kr) − ki h` (kr) j` (kr0 )r02 dr0 U (r0 )ψ` (kr0 ). (7.28)
r→∞ 0

h` (ρ)+h∗
` (ρ) ei(ρ−`π/2)
Since j` (ρ) = 2 , and h` (ρ) −→ iρ , for the asymptotic wave function we get
ρ→∞

(+) 1 ∗ 1
ψ` (kr) → h (kr) + h` (kr)[1 + 2kiT` (k)]
2 ` 2
1 e−i(ρ−`π/2) e+i(ρ−`π/2)
→− { − S` (k)},
2i ρ ρ
CHAPTER 7. SCATTERING-CONTINUED 163

with ρ = kr. Where the definitions


Z ∞
(+)
T` (k) = − j` (kr)U (r)ψ` (kr)r2 dr (7.29)
0

and
S` (k) = 1 + 2kiT` (k) (7.30)
have been introduced. (S` (k) is often called the partial wave S-matrix.)
Note that for zero potential (U = 0) we get T → 0, and S` → 1; also, the wave function returns
to the form obtained from the plane wave decomposition alone; namely,

(+) 1 e−i(ρ−`π/2) e+i(ρ−`π/2)


ψ` (kr) → j` (kr) −→ − {1 −1 }.
ρ→∞ 2i ρ ρ

Here we see that the free case (no scattering) consists of incoming and outgoing spherical waves of
equal amplitude 1. Hence the radial flux in and out are equal in magnitude.
For elastic scattering no flux should be lost when there is an Hermitian potential and thus
we require for elastic scattering that the amplitude of the incoming and outgoing spherical waves
continue to have unit magnitude even for a potential present.

... | S` (k) |= 1

(for elastic scattering). We thus have S` (k)S`∗ (k) = 1, or by introducing real quantities called phase
shifts δ` (k), we have ∗
S` (k) = e2iδ` (k) ; S`∗ (k) = e−2iδ` (k) = e−2iδ` (k) . (7.31)
Here we see that real phase shifts are required for elastic scattering; i.e., for no flux lost. We will
see later that δ` (k) is properly named since it gives a shift in the phase of the outgoing wave (but
keeps the amplitude of unit magnitude.) With

S` (k) = e2iδ` (7.32)

it follows that
S` (k) − 1 e2iδ` − 1 eiδ` (k) sin δ` (k)
T` (k) = = =
2ki 2ki k
or

eiδ` sin δ` (k)
Z
(+)
=− j` (kr)U (r)ψ` (kr)r2 dr
k 0

2m ∞
Z
(+)
=− 2 j` (kr)V (r)ψ` (kr)r2 dr. (7.33)
h̄ 0
This relates the partial wave amplitude and the associated phase shifts to the Bessel transform of
the potential times the short distance wave function. ( Note we really should use the reduced mass
instead of m.)

7.4.5 Solutions of the Radial Differential Equation


Our approach has been simply to introduce the partial wave decomposition of the integral equation,
which already incorporates the outgoing wave boundary condition. Another way to deduce the form
of the asymptotic wave function is to begin with the radial equation

h̄2 1 ∂ 2 ∂ h̄2 `(` + 1) h̄2 2


[− r + + V (r)]ψ ` (kr) = k ψ` (kr),
2µ r2 ∂r ∂r 2µ r2 2µ

where we replace the notation used for the bound state case R` (ρ) by ψ` (kr). For distances outside
of the range of V (r) (a remark that can not be satisfied for a Coulomb potential), the above equation
has the solutions j` (kr) and n` (kr)

ψ` (kr) −→ A` j` (kr) + iB` n` (kr). (7.34)


r>a
CHAPTER 7. SCATTERING-CONTINUED 164

The choice of A` is based on our need to reconstitute the plane wave and for
r
(+) 2 X ` (+) ∗
ψ~ (r) ≡ i ψ` (kr)Y`m (r̂)Y`m (k̂), (7.35)
k π
the correct choice is A` = 1 + B` . The reason for this choice is that the plane wave part (1 j` (kr))
is obtained and an outgoing spherical wave part (B` h` (kr)) is also generated.

ψ` (kr) −→ j` (kr) + B` h` (kr)


r>a

h∗` (kr) h` (kr)


−→ + {1 + 2B` } (7.36)
r>a 2 2
- as before. The last arrangement leads us again to S` (k) = e2iδ` (k) = 1 + 2B` for the previously
discussed reason that flux is conserved. Hence, we have 2B` = 2kiT` (k) = 2ieiδ
` sin δ` . The boundary
conditions have been imposed properly on the solutions of the differential equations.

Cross Section
The cross section for scattering is determined by the amplitude of the scattered wave. Recall that
ikr
(+) 1 i~
k·~
r ~k 0 , ~k) e }
ψ~ (r) −→ {e + f ( (7.37)
k (2π)3/2 r

defined the scattering amplitude f (~k 0 , ~k) and the elastic scattering cross section was obtained as
dσ 2
dΩ =| f | .
Using the partial wave decomposition and its asymptotic behavior we see that
(+)
ψ` (kr) −→ j` (kr) + ikh` (kr)T` (k)
Z ∞
2µ (+)
T` (k) = − 2 j` (kr0 )V (r0 )ψ` (kr0 )r02 dr0 . (7.38)
h̄ 0
Therefore
~ r
(+) eik·~r 2X ` ∗
Ψ~ (r) −→ + i ik h` (kr) T` (k) Y`m (r̂)Y`m (k̂)
k (2π)3/2 π
`m

1 ~
X ei(kr−`π/2) ∗
−→ [eik·~r + 4π i` k T` (k) Y`m (r̂) Y`m (k̂).
(2π)3/2 `m
kr

– ≡ k −1
Thus, after using r̂ ≡ k̂ 0 and λ
X
f (~k 0 , ~k) = 4π T` (k)Y`m (k̂ 0 )Y`m

(k̂)
`m
X

= 4πλ [eiδ` (k) sin δ` (k)]Y`m (k̂ 0 )Y`m

(k̂)
`m
X

=λ [eiδ` (k) sin δ` (k)](2` + 1)P` (k̂ 0 · k̂)
`
–X

= [1 − S` (k)](2` + 1)P` (k̂ 0 · k̂) (7.39)
2
`

The differential cross section for elastic scattering is therefore given by



dσ –2 |
X
=| f |2 = λ eiδ` (k) sin δ` (k)(2` + 1)P` (cos θ) |2 (7.40)
dΩ 0
X
–2
= λ [eiδ` (k) sin δ` (k)e−iδ`0 (k) sin δ`0 (k)]
``0
(2` + 1)(2`0 + 1)P` (cos θ)P`0 (cos θ).
CHAPTER 7. SCATTERING-CONTINUED 165

The total elastic scattering cross section, obtained by integrating over all solid angle has the
simple structure Z X
σsc = dΩ | f |2 = λ –2 sin2 δ` (k)(2` + 1)4π
`
X
–2
σsc = 4πλ (2` + 1) sin2 δ` (k), (7.41)
`

since Z

dΩ P` (cos θ) P`0 (cos θ) = δ``0 . (7.42)
2` + 1

Note that for no scattering U = 0, and δ` = 0, and both σsc → 0 and dΩ → 0 as they should.
Another convenient way of writing the total elastic scattering cross section is in terms of S` (k)
X
σsc = πλ–2 (2` + 1) | 1 − S` (k) |2 , (7.43)
`

which vanishes for S` (k) → 1. We used

| 1 − S` (k) |2 =| eiδ sin δ` 2i |2 = 4 sin2 δ` .

7.4.6 Optical Theorem–Partial Waves


The optical theorem can be used to first check the above results and then extend them to include
absorption or reaction processes. First let us check the above elastic scattering results. The optical
theorem is

σT OT = + Im felastic (θ = 0◦ ). (7.44)
k
For elastic scattering the total cross section consists of elastic scattering alone (σT OT ≡ σSC ) and
we have X
σSC = +4πλ – Im felastic (θ = 0◦ ) = +4πλ– Im –eiδ` sin δ` (2` + 1)
λ
`

since P` (k̂ · k̂) = P` (1) = 1 for all `. Thus


X
–2
σSC = 4πλ sin2 δ` (2` + 1), (7.45)
`

–2 (2` + 1). i
which checks the above result. Note for a give `-value: σ`sc ≤ 4πλ

Generalized to absorption
In general, the outgoing flux might be diminished by various reactions that remove flux from the
elastic channel. In that case | S` |< 1 absorption (or reactions).
With this condition, we can introduce complex phase shifts

δ` (k) = δ`R + iδ`I

and R I
S` = e2iδ` = e2iδ` e−2δ`
I
| S` |= e−2δ` < 1 for δ`I > 0
(if δ`I < 0, | S` |> 1 and we would have creation or production of flux.)
The optical theorem for the total cross section then gives us

σT OT = σsc + σab = Im felastic (θ = 0◦ ), (7.46)
k
i See zones picture later.
CHAPTER 7. SCATTERING-CONTINUED 166

where the total cross section now consists of the elastic scattering plus the absorption (and/or
reactions, inelastic collisions · · ·) cross section to account for lost flux. Remembering that δ` is now
complex, we get
– Im [ 1 iλ
X
σT OT = 4πλ – [1 − S` (k)](2` + 1)] (7.47)
2
`

–2 1 Im
X
= 4π λ i(1 − ReS` − iImS` )(2` + 1)
2
`
X
–2
σT OT = 2πλ (1 − ReS` )(2` + 1)
`
= σsc + σab .
Here σab represent everything that can occur aside from elastic collisions.
We can hence derive an expression for the σab using

σab = σT OT − σsc (7.48)


X
–2
= πλ (2` + 1){2(1 − Re(S` ))− | 1 − S` |2 }
`
X
= πλ (2` + 1){2(1 − Re(S` )) − (1 − S` )(1 − S`∗ )}
–2
`
X
= πλ –2 (2` + 1){2(1 − Re S` ) − 1 + S`∗ + S` − | S` |2 }
`
X
= πλ –2 (2` + 1){1 − 2Re S` + 2Re S` − | S` |2 }
`
X
–2
= πλ (2` + 1){1− | S` |2 }.
`

Note if | S` |= 1, we get σab = 0 corresponding to no flux lost and to elastic scattering only!

7.4.7 Examples
Absorbing Sphere
Let us consider what happens if all of a set of lower partial waves less than some `max are totally
absorbed.

δ`I = ∞ for ` < `max ∼


= kR
... S` (k) = e2iδ` e−2δ` = 0 for ` < `max and δ = 0 for ` ≥ `max
R I

X−1
`max
–2
σT OT = 2πλ –2 [kR]2 = 2πR2
(2` + 1) = 2πλ
`

X−1
`max
–2
σabs = σnel = πλ –2 (kR)2
(2` + 1) = πλ
0

σel = πR2
*We take nearest integer and assume that `max is quite large; i.e. a high energy approximation.)
CHAPTER 7. SCATTERING-CONTINUED 167

7.4.8 Meaning of Phase Shifts


We have introduced a phase change in the outgoing flux S` = e2iδ` , which for elastic scattering
causes no change in the amplitude or no flux lost | S` |= 1. The resulting asymptotic wave function
is for r > a, where a = range of potential.
(+)
ψ` (kr) −→ j` (kr) + ikh` (kr)T` (k)
eiδ` sin δ` 2µ ∞
Z
(+)
T` (k) ≡ =− 2 j` (kr)V (r)ψ` (kr)r2 dr. (7.49)
k h̄ 0
Using the asymptotic forms of j` and h`
(+)
krψ` (kr) −→ sin(kr − `π/2) + ei(ρ−`π/2) kT` (k)

−→ sin(kr − `π/2)[1 + ikT` (k)]


+ cos(kr − `π/2)kT` )
−→ eiδ [sin(kr − `π/2) cos δ` + cos(kr − `π/2) sin δ` ]
... krψ` (kr) −→ eiδ sin(kr − `π/2 + δ` ).
(+)
(7.50)
r→∞

Note that the phase −`π/2 originates from the centrifugal barrier repulsion, from which it is clear
that a repulsion yields a negative phase shift. The centrifugal barrier is incorporated into the free
spherical Bessel functions. The “phase shift” δ` (k), which is a function of orbital angular momentum
` and energy, is a shift in phase relative to the free asymptotic form sin(kr − `π/2), as shown in the
following figure

ã-i ∆ k r Ψ
1

0.5

r
2 4 6 8 10

-0.5

-1

(+)
Figure 7.11: A plot of the asymptotic wave function kre−iδ` ψ` (kr) versus r. The solid line is the
original krj` (kr), and the dashed curve the phase-shifted asymptotic wave function sin(kr−`π/2+δ` ),
where δ` = π/4 is attractive.

The sign conventions are therefore that for an attraction ( V < 0) the phase shift is positive
δ` > 0, whereas for a repulsion V > 0 the phase shift is negative δ` < 0.
The phase shift can be shifted to another quadrant without changing any observable since under
the change δ` → δ` + 2nπ, we get ei(δ+2nπ) sin(δ + 2nπ) = eiδ sin δ.

Levinson’s Theorem
Let me state Levinson’s theorem, without proof. This theorem fixes the phase shift at zero energy
δ` (0) assuming that we set the E → ∞ phase shift as zero δ` (∞) = 0. The difference

δ` (k = 0) − δ` (k = ∞) = n`B π
where n`B is the number of bound states with angular momentum ` supported by the potential.
CHAPTER 7. SCATTERING-CONTINUED 168

Hard Core
Consider an infinite repulsion within a range a. V = ∞; r ≤ a and V = 0; r > a. The wave function
must vanish at r ≤ a. Therefore, since we can use the asymptotic form for r ≥ a
(+)
ψ` (kr) −→ j` (kr) + ikT` (k)h` (kr)

Since ψ` = 0 inside the hard core r ≤ a

j` (ka)
kT` (k) = −
ih` (ka)
or
sin δ` tan δ` −j` (ka)
eiδ` sin δ` = = =
cos δ` − i sin δ` 1 − i tan δ` ij` (ka) − n` (ka)
j` (ka)/n` (ka)
=
1 − ij` (ka)/n` (ka)
j` (ka)
... tan δ` (ka) =
n` (ka)
Also:
X (2` + 1)j 2 (ka)
–2
σ = 4πλ `
.
j`2 (ka) + n2` (ka)
For ` = 0 r
sin ka . 2m 2
tan δ0 = − .. δ0 = −ka = − Ea
cos ka h̄2
For low energies, only get S-waves and
dσ –2 sin2 δ0 ⇒ λ
–2 a2 k 2 = a2

dΩ
σsc = 4πa2
We can also show that for low energy j` (ka) → (ka)` /(2` + 1)!! and n` (ka) → −(2`−1)!! (ka)`+1
and
hence at low energies
(ka)2`+1
tan δ` (k) ≈ δ` (k) = ,
(2` + 1)!!(2` − 1)!!
which falls off rapidly with `... at low k only S-waves contribute. Note the double factorial (2n+1)!! ≡
1 · 3 · 5 · · · n.

(We could look at effect of P-waves on dΩ , etc.)
CHAPTER 7. SCATTERING-CONTINUED 169

7.5 Resonances
To understand resonances one important model is an attractive, 3-D Square Well. (Ref. Merzbacher
and Blatt & Weiskopf) V (r) = −V0 for r ≤ a and V (r) = 0 for r > a.

VHeffectiveL

400

200

r
1 2 3 4

-200

Figure 7.12: An attractive square well of depth −V0 for r ≤ a, plus a centrifugal barrier. This is for
` = 1.

In addition to this attractive well, there is a repulsive centrifugal barrier `(` + 1)/r2 . This model
also provides a more general result, since it uses a boundary condition to stipulate the internal
dynamics. Indeed, this was used by E. Wigner to formulate a general boundary condition theory of
reactions. q
(+)
The solution of the radial equation is simply ψ` (kr) = j` (α r) for r ≤ a where α ≡ k 2 + 2m h̄2
| V0 |.
At r = a, we have a logarithmic derivative
dψ`
dr j`0 (αa)
γ` ≡ |r=a = α , (7.51)
ψ` j` (αa)

which is a known function of energy E, potential depth V0 , and potential range a. Here we denote
j`0 = dρ
d
j` (ρ). For r ≥ a, the asymptotic wave function is given as

(+)
ψ` (kr) → j` (kr) + ikT` (k)h` (kr). (7.52)

Equating the inside and outside the well logarithmic derivatives, to assure continuous probability
and current, at r = a, we have

kj`0 (ka) + ik 2 T` (k)h0` (ka)


γ` =
j` (ka) + ikT` (k)h` (ka)

γ` j` (ka) − kj 0 (ka) tan δ`


... kT` (k) = eiδ` sin δ` = =
i[kh0` (ka) − γ` h` (ka)] 1 − i tan δ`
−γ` j` (ka) + kj`0 (ka)
=
i[−kj`0 (ka) + γ` j` (ka) − ik n0` (ka) + γ` in` (ka)]
kj`0 (ka) − γ` j` (ka) j` (ka)
... tan δ` (k) = 0 =⇒ → “hard core limit.” (7.53)
kn` (ka) − γ` n` (ka) γ` →∞ n` (ka)
Note that γ` → ∞ when the wave function vanishes at r = a and for that reason the result reduces
to a hard core answer. The wave function inside r = a does not have to vanish to get this limit.
Another way of expressing this `th partial wave amplitude is:
1 + i tan δ` 2i tan δ`
e2iδ` = S` (k) = 1 + 2ikT` (k) = =1+ (7.54)
1 − i tan δ` 1 − i tan δ`
D` + iN`
e2iδ` = ;
D` − iN`
with
N` = kj`0 (ka) − γ` j` (ka)
D` = kn0` (ka) − γ` n` (ka)
CHAPTER 7. SCATTERING-CONTINUED 170

and
N`
tan δ` =
.
D`
Another clever separation involves pulling out the “hard core” part. The first step is to define
the “hard core” (or γ` → ∞ limit) phase shift ζ` by
j` (ka)
tan ζ` ≡ .
n` (ka)
and then consider
tan δ` − tan ζ`
tan(δ` − ζ` ) = (7.55)
1 + tan δ` tan ζ`
N` j` (ka)
D` − n` (ka)
= N` j` (ka)
1+ D` n` (ka)
N` n ` − j ` D `
=
D` n` + j` N`
[kj`0 − γ` j` ] n` − [j` kn0` − γ` n` ]
=
[j` kn0` − γ` n` ] n` + j` [kj`0 − γ` j` ]
kj`0 n` − kj` n0`
= .
k(n` n0` + j` j`0 ) − γ` (n2` + j`2 )

Now we know that the Wronskian j`0 n` − j` n0` = − (ka)


1
2 , so

−1/a
tan(δ` − ζ` ) = (7.56)
ka(n` n0` + j` j`0 ) − γ` a(n2` + j`2 )
1 1
ka n2` +j`2
= ka(j` j`0 +n` n0` )
γ` a − n2` +j`2
s`
=
β` − ∆`
where we have defined
1 1
s` = (7.57)
ka n2` (ka) + j`2 (ka)
β` = γ` a
ka(j` (ka) j`0 (ka) + n` (ka) n0` (ka))
∆` = .
n2` (ka) + j`2 (ka)
Also the combination
j`0 (ka) + in0` (ka) h0 (ka)
∆` + is` ≡ ak = ak `
j` (ka) + in` (ka) h` (ka)
is often used.
These are the characteristic functions that will be used to isolate resonances and generate the
general form of the S-matrix. Note that s` and ∆` are real and that we introduce s` = kav` and
have
∆` = (ka)3 v` (j` j`0 + n` n0` ).
Now form
1 + i tan(δ` − ζ` )
e2i(δ` −ζ` ) = (7.58)
1 − i tan(δ` − ζ` )
β` − ∆` + is`
=
β` − ∆` − is`
or
β` − ∆` + is`
S` = e2iδ` = e2iζ` . (7.59)
β` − ∆` − is`
CHAPTER 7. SCATTERING-CONTINUED 171

If the log derivative part s` is real, then the above S-matrix is unitary S`∗ S` = 1, for each `.
Note s` (k) has the limit properties

s` (k) ∝ (ak)2`+1 ; ka → 0

and
s` (k) ∝ (ak); ka → ∞
Also s` , ∆` and ζ` are all slowly varying functions of `, k and the range a.
It is the logarithmic derivative β` ≡ aγ` that specifies the interior dynamics. Hence, the above
result applies to any general finite range potential once β` is provided. Once β` (k) is specified, the
cross sections are determined. j
For a square well potential β` is given by

d (+) (+) j 0 (αa)


β` = aγ` = a( ψ` )/ψ` |r=a = aα ` ,
dr j` (αa)
q
α = k 2 + 2m h̄2
| V0 |. The function β` , which depends on `, E, a and V0 is a real, unitless, monoton-
ically decreasing function of energy.
To get the scattering amplitude, we express eiδ` sin δ` as
S` − 1
kT` (k) = eiδ` sin δ` =
2i
e2iζ` 2is` 1
= {1 + }−
2i β` − ∆` − is` 2i
s` 1
= e2iζ` { + (1 − e−2iζ` )}
β` − ∆` − is` 2i
s`
... iδ` 2iζ`
e sin δ` = e { + e−iζ` sin ζ` }
β` − ∆` − is`
Thus the amplitude has been separated into a resonant plus a nonresonant or background term.
To see the resonant term more clearly note that:

β` − ∆` + is` 1 + i tan(δ` − ζ` )
e2i(δ` −ζ` ) = =
β` − ∆` − is` 1 − i tan(δ` − ζ` )

and
s`
tan(δ` − ζ` ) =
β` − ∆`
For the phase shift we obtain:
s`
δ` = ζ` + tan−1 .
β ` − ∆`
Assume a linear dependence on energy for β`

β` = a` − Ec` ζ>0
s` Γ`
and define the width function Γ` by c` ≡ 2 >0

s` /c` Γ`
δ` = ζ` + tan−1 a` −∆`
= ζ` + tan−1
C` − E 2(ER` − E)

Therefore, for a linear dependence of β` on energy E, the resonant form is obtained as


Γ`
δ` = ζ` + tan−1
2(ER` − E)
j When generalized to multichannel problems, the above result that S depends on γ only, becomes the basis of
` `
general nuclear (and atomic) reactions theories as developed by Wigner - see Blatt and Weiskopf’s discussion.
CHAPTER 7. SCATTERING-CONTINUED 172

with the “width” given (in MeV) by the positive quantity


2s` v`
Γ` = = 2ka > 0.
c` c`
Note Γ` is determined by the penetration factor v` (which depends on the centrifugal barrier) and
by c` ∝ − dβ
dE |0 , the rate of change of the logarithmic derivative.
`

The “resonant energy” is given by


a ` − ∆`
ER` = ,
c`

where a` ' β` (ER ), and c` ' − dβ


dE |ER are determined by the dynamics in the interior region and
`

∆` is the “shift function,” which depends on the centrifugal barrier.


Evaluating the above results at E = ER` ∓  leads to

δ` = ζ` + π/2 ∓ 

and hence
Γ`
δ` − ζ` = tan−1
2(ER` − E)
yields the following resonant phase shift curve for a single resonance at ER` . [We also use δ` −ζ` αk 2`+1
and Γ` αk 2`+1 at low energy.] The slope at E = ER` is given by
∆L
3
2.5
2
1.5
1
0.5
ã
20 40 60 80 100

Figure 7.13: The phase shift δ` − ζ` versus energy plotted near a resonance for two different widths.
The steeper slope (dashed curve) corresponds to a larger time delay and a smaller width. Note the
phase shift is π/2 at the resonance energy ER .

d
(δ` − ζ` ) |E=ER` = 1/Γ` /2;
dE
therefore a wide (Γ` large) and a narrow resonance (Γ` small) give phase shifts as shown in the
figure.
It is usually assumed that ER` and Γ` vary slowly with energy and hence can be treated as
constants in the resonance region E near ER` . At low energies, one needs to recall that δ` , ζ` and
Γ` have the “threshold” behavior of ∼ k 2`+1 . Using the above linearized logarithmic derivative
(β` ∼
= a` − Ec` ), the result for the scattering amplitude is
Γ` | 2
eiδ` sin δ` = e2iζ` { + e−iζ` sin ζ` }.
ER` − E − iΓ` /2
If we ignore the hard core phase shift ζ` , then the total scattering cross section is

–2
X Γ2` /4
σSC = 4πλ (2` + 1) Γ2`
,
` (ER` − E)2 + 4

which is the famous Breit-Wigner resonance form with ER` and Γ` constant near E = ER` )
At E = ER` , the `th term dominates and assumes its maximum value; here we assume nonover-
lapping resonance so only one isolated term is picked out at its resonant energy. Then σs ' σSC`
with
max
σSC,` –2 (2` + 1) at E = ER` ,
= 4πλ
CHAPTER 7. SCATTERING-CONTINUED 173

1 max
σSC,` = σ at E = ER` + Γ` /2.
2 SC,`
Hence Γ gives the full resonance width. k
Near this resonance the differential cross section is given by

dσ –2 (2` + 1)2 P`2 (cos θ) Γ2` /4


'λ .
dΩ (ER` − E)2 + Γ2` /4

Therefore, from the differential cross section one can determine the `-value associated with the
elastic scattering resonance. (Useful for assignments of unknown quantum numbers in various situ-
ations such as observation of nuclear, meson or baryon resonances.)
The scattering amplitude can also be given for the resonant case as
X
f (~k 0 /~k) = λ
– (2` + 1)eiδ` sin δ` P` (cos θ)
`


X Γ` /2
=λ (2` + 1){ }P` (cos θ)
`
(ER` − E)−1 Γ2`
for (ζ → 0). We note that the real and imaginary parts of f for θ = 0 :
X –Γ` /2
λ
Refk (θ = 0) = Γ2`
(ER` − E)(2` + 1)
` (ER` − E)2 + 4
and
X –Γ2` /4(2` + 1)
λ
Imfk (θ = 0) = > 0.
(ER` − E)2 + Γ2` /4
`

4π (2`+1)Γ2 /4
= 4πλ2
P
The optical theorem is here verified, since σsc = k Imf (0) Γ2
.
(E−RR` )2 + `
4
Also note that as functions of energy these yield
The Imfk (0) has its maximum, while Refk (0) has its zero at the resonance energy E = RR` .
The expression for e2iδ` is also of general interest. Again ignoring ζ` , we obtain a simple result

ER` − E + i Γ2`
S` (k) =
ER` − E − i Γ2`

which shows that S` has a simple pole in the complex energy plane at E = ER` − iΓ` /2. (More
about this later).
Also, we have S` (k)S`∗ (k) = 1 provided ER` and Γ` are real. Making ER` or Γ` complex would
introduce absorption and lead to a nonzero absorption cross-section σab . See H.W. problem on case
of | S` |< 1.

7.6 Time Delay and Resonance


To understand the physical significance of a simple potential resonance, it is worth considering the
following simplified version of the Wigner-Eisenbud demonstration of a time delay near a resonance.
One learns that a resonance corresponds to the particles spending some time together (i.e. resonat-
ing), before they part company and contribute to the outgoing radial flux. We show below that the
time delay is given by
1 ∂ h̄
tD = δ` (k) =⇒ = 2τ1/2 ;
vg ∂k E⇒ER` (Γ` /2)

Here τ1/2 = h̄/Γ` . This time delay corresponds to the time the two particles live together - τ1/2 is
often called the “lifetime” of the resonance.
k see later where it is shown that ẼR` ≤ ER` .
CHAPTER 7. SCATTERING-CONTINUED 174

Proof:
We need to change only one step in our earlier version of wave packet scattering; namely, the step
involving factoring
fk (θφ)
out of the scattered wave part

d~kφ(~k − ~k0 ) eikr −iω(k)t


Z
ψsc (~rt) → 3/2
fk (θφ) e
(2π) r

–(2` + 1)eiδ` (k) sin δ` (k)P` (cos θ), where cos θ refers to the ~r · ẑ
A simple resonance gives us fk (θφ) = λ
angles. Expanding δ` (k) and keeping the higher order term in the phase, we get

δ` (k) = δ` (k0 ) + (k − k0 )δ`0 (k0 )

and 0
fk (θφ) = ei(k−k0 )δ` (k0 ) fk0 (θφ)
. Inserting into ψsc , we get

d~kφ(~k − ~k0 ) ik(r+δ`0 (k0 ) −iω(k)t


Z
0
ψsc (~rt) → fk0 (θφ)/r e−ik0 δ` (k0 ) e e
(2π)3/2
0
Thus the previous steps apply, except we replace r by r + δ`0 (k0 ). (The phase eik0 δ` (k0 ) makes no
difference). Thus we can repeat the previous steps and find
~
Ψ(+) (~r, t) = eiω0 t {A(~r − ~v0 t)eik0 ·~r

fk0 (θφ) 2iω0 t −ik0 δ`0 (k)


+ e e A(r + δ`0 (k0 ) − v0 t)}
r
The extra term in the outgoing scattered wave packet corresponds to a time delay:
1 0
r + δ`0 (k0 ) − v0 t = r − v0 (t − δ (k0 ))
v0 `
= r − v0 (t − tD )
The outgoing spherical wave packet is held back l a delay time
1 ∂
tD = δ` (k) |k0
v0 ∂k
2m ∂ ∂
=δ` (k) |k0 = h̄ δ` (E) |E0
h̄ ∂k 2 ∂E
At the resonance E0 = ER and we use our earlier result to deduce the time delay to be determined
by Γ` :
2h̄
tD |ER = QED .
Γ`
Wigner also showed that the requirement of causality (i.e. the outgoing wave develops only after
contact is made) leads to the “Wigner condition”

1 ∂ 2a
−tD = − δ` (k) |k0 <
v ∂k v0
a = range of V, or

δ` (k) > −2a.
∂k
l There can be an advanced time, as for a pure S-wave hard core, but no so advanced as to violate causality
CHAPTER 7. SCATTERING-CONTINUED 175

7.6.1 Wave Functions and Resonances


Another way of visualizing the phenomena of resonance is to examine the wave function. At a
resonance the interior wave function is large corresponding to a high probability of the particles
spending an appreciable time together - as obtained in the previous time-development description.
The basic physics involved in a resonance is essentially that given by the classical problem of a heavy
rope driven by a lighter one. q
This analogy is based on the observation that the internal momentum α = k 2 + 2m h̄2
| V0 | is
larger than k; the corresponding internal wave length is thus smaller than the exterior wave length
–inside < λ
α > k and λ –exterior .

We therefore have the problem of matching a high wave length exterior to a low wave length interior
wave function with ψ and ψ 0 kept continuous. (This is essentially a question of impedance matching).
For an off-resonance case, the wave function looks like:
off-resonance – figure to be inserted
| ψ |2 for r < a small. ... ψ exterior behaves as if scattered from a hard core of radius a.
Near resonance—-figure to be inserted
On Resonance—-figure to be inserted
| ψ |2 for r < a largest!
CHAPTER 7. SCATTERING-CONTINUED 176

7.6.2 Resonances - Example of Separable Potential


As a further example of the dynamics involved in the phenomena of a resonance, let us consider
the separable potential case, which was also discussed for the bound state problem in earlier Home
Work problems.
We first generalize the definition of a nonlocal potential to permit it to act separately in each
partial wave ` X
(~r | V | ~r 0 ) = (r | v` | r0 )Y`m (r̂)Y`m

(r̂0 ), (7.60)
`m

which defines a nonlocal central potential in each partial wave `. Although this form is nonlocal, it
~ V ] = 0.
still yields a central potential since [L,
A separable potential is a special case of a nonlocal interaction for which the function v` of the
relative coordinates r =| ~r |=| ~r1 − ~r2 | and r0 =| ~r 0 |=| r10 − r20 | is assumed to be a product

h̄2
(r | v` | r0 ) = − g` (r)g` (r0 ), (7.61)
2m
which defines an attractive separable potential in the `th partial wave, with Mev/F 3 ( energy per
volume) units for (r | v` | r0 ).
The momentum space expression involves finding < ~k | V | ~k 0 > to be
X
< ~k | V | ~k 0 >= (k | v` | k 0 )Y`m (k̂)Y`m

(k̂ 0 )
`m

2
with (k | v` | k 0 ) = − 2m
h̄ 2 0
π g` (k)g` (k ). The function g` (k) is simply the Bessel transform
Z ∞
g` (k) = dr r2 j` (kr) g` (r),
0

and correspondingly m Z ∞
2
g` (r) = dk k 2 g` (k) j` (kr).
π 0
α
For example, for the S-wave ` = 0 case we can take g0 (k) = k2 +a2 which gives

g0 (r) = αe−ar /r,

where a is an inverse range.


Using this decomposition of the potential, along with the partial wave decompositions presented
(+)
earlier for the Green’s function, plane wave, and wave function ψ~ , one obtains
k
Z ∞
2m (+)
ψ (+) (r) = j` (kr) + (r | G` (k 2 + i) | r0 )(r0 | v` | r00 )ψ` (kr00 )r002 dr00 r02 dr0 .
h̄2 0

For the separable potential this becomes


Z ∞
(+) (+)
ψ` (r) = j` (kr) − (r | G` (k 2 + i) | r0 )g` (r0 )r02 dr0 < g` | ψ` >,
0

with the definition Z ∞


(+) (+)
< g` | ψ` >≡ g` (r) ψ` (r)r2 dr.
0
Resorting to some formal notation, we have
(+) (+)
| ψ` >=| j` > −G` | g` >< g` | ψ` >,
(+) (+)
< g` | ψ` >=< g` | j` > − < g` | G` | g` >< g` | ψ` >,
0 0 R∞
m From < ~k | ~k >= δ(~k − ~k ), one can show that
2 2
π
0 dr r j` (kr) j` (k r) = δ(k − k0 )/k2 .
0
CHAPTER 7. SCATTERING-CONTINUED 177

(+) g` (k)
< g` | ψ` >= R ∞ dk0 k02 g`2 (k0 )
2
1− π 0 k02 −k2 −i
To get the scattering amplitude, we need to evaluate
Z ∞
2m (+)
eiδ` (k) sin δ` (k) = −k j` (kr)(r | v` | r0 ) 2 ψ` (r0 )r02 dr0 r2 dr.
0 h̄
(+)
= +kg` (k) < g` | ψ` >
or
kg`2 (k) kg`2 (k)
eiδ` (k) sin δ` (k) = R ∞ dk0 k02 g`2 (k0 ) = D(+) (k 2 ) .
2
1− π 0 02 2
k −k −i

We introduce the denominator function (also called a ‘Fredholm determinant” or a “Jost function”)
defined by
2 ∞ dk 0 k 02 g`2 (k 0 )
Z
D(k 2 + i) ≡ D(+) (k 2 ) ≡ D(+) (E) ≡ 1 −
π 0 k 02 − k 2 − i
Z ∞ 0 0 2 0
2 dk k g` (k )
=1− P − ikg`2 (k)
π 0 k 02 − k 2
= D(k 2 ) − ikg`2 (k)
Here P denotes the principal value integral, and D(k 2 ) denotes the principal value integral. The last
step separates out the principal value integral and enables us to write (using eiδ sin δ = tan δ/(1 −
i tan δ))
λkg`2 (k) kg 2 (k)
tan δ` = ∞ dk 0 k 02 g 2 (k 0 )
≡ ` 2
1 − 2P
R
02
`
2
D(k )
π 0 k −k
This solution to the scattering problem can be used to display the analytic properties of the scattering
amplitude and to show how resonances arise.

Analytic properties
First, let us consider the role of the bound state condition. The existence of a bound state requires
that
2 ∞ dk 0 k 02 g`2 (k 0 )
Z
2 = 1 or that D(+) (−EB ) = 0.
π 0 k 02 + kB
(see earlier Homework problem). Hence, the scattering amplitude viewed as a function in the complex
energy plane is seen to have a pole at the bound state energy E = −EB . This followed earlier in
general from the Low equation T (E) = V + V (1/(E − H))V.
Other singularities of eiδ` sin δ` in the complex energy plane can now be identified. A branch cut
arises from the integration over k 0 from the zero energy threshold to + ∞. Potential cuts are the
complex poles or cuts in g` (k).
The resonance pole follows from the expression for tan δ` . The phase shift δ` passes up through
π/2 if the principal value function
2
D(kR )=0
or ∞
dk 0 k 02 g`2 (k 0 )
Z
2
P 2 = 1.
π 0 k 02 − kR`
2
h̄ 2
This gives the condition for a resonance to occur at E`R = 2m kR` . Corresponding to this singularity
iδ`
in tan δ` , the amplitude e sin δ` has a pole at E = ER` − iΓR` , where we can see that the width is
determined from
Γ`
tan δ` '
2(ER − E)
in the vicinity of the resonance. Note:
kg`2 (k)
tan δ` = 2
R∞ 2 1 1
πP 0
dk k g` (k 0 )[ k02 −k
0 02
2 − k02 −k2 ]
R`
CHAPTER 7. SCATTERING-CONTINUED 178

R∞ dk0 k02 g`2 (k0 )


kR g`2 (kR )/ π2 P 0 (k02 −kR` 2 )2

= ,
2 − k2 )
(kR`
using the fact that kg` (k) and the integral are slowly varying functions near ER . The above results
permit us to identify the width as
Z ∞ 0 02 2 0
h̄2 2 2 dk k g` (k )
ΓR` = k`R g` (kR` )/[ P 2 )2 ]
2m π 0 (k 02 − kR`

Several important models make use of this separability of the interaction - for example the Lee
model and Chew-Low models in field theory.

Delta Shell potential


The above discussion reduces to a ‘Delta-shell” potential if we make the special choice (see H.W.)
1
g` (r) = α δ(r − r0 )
r2
for all `. This choice yields

h̄2 1
(~r | V | ~r0 ) = −α2 δ(r − r0 )δ(~r − ~r 0 )
2m r02

Delta-shell or Hard Shell which is a “local” potential.

h̄2 α2
V (r) = − δ(r − r0 )
2m r02
Then we get g` (k) = j` (kr0 ) and

2 ∞ dk 0 k 02 j`2 (k 0 r0 )
Z
= +kij` (kr0 )h` (kr0 )
π 0 k 02 − k 2 − i

(see previous evaluation of (r | G` (k 2 + i) | r0 ).

The complex k-plane


Another topic is the description of the analytic properties not only in the complex energy plane, but
also in the complex k-plane. The relationship is

h̄2 2
E= k =| E | e2iφ complex E
2m
k =| k | eiφ complex k
Hence a rotation in the k-plane of 180◦ brings one to the negative side of the branch cut. For
example, we have

... upper (Im k > 0) half of k-plane corresponds to the first (“physical”) Riemann sheet in the
Energy plane. The lower (Imk < 0) half corresponds to the second Riemann sheet obtained by
continuation through the branch cut. The discontinuity across the cut is see to be related to the
values of the scattering amplitude.
CHAPTER 7. SCATTERING-CONTINUED 179

It can be shown that causality (scattering occurs only after contact) requires that all other
singularities of the scattering amplitude eiδ` sin δ` in the complex k-plane occur in the lower 21 -plane
(Imk < 0). These singularities include the resonance poles at

k = kR` eiφR with3π/2 < φR < 2π.

and the potential poles (at k = −ia) or potential cuts which arise from the functions g` (k). (These
potential cuts involve the range of interaction due to meson exchanges.)
It should be noted that none of these fancy ways of recording the occurrence of bound states
and resonances is really needed for our purpose. Analyticity is simply a result of the well-behaved
nature of the Schrödinger equation for small changes of the energy, even when extended to complex
values. It is only at bound states and resonances that singular-type behavior (poles) occur. (The
continuum states are unbounded and are here expressed by the branch cut.) As long as a well-
behaved V (r) is given, one can solve the Schrödinger equation without a discussion of analytic
properties. Aside from the extra insight gained, the only reason for discussing analyticity properties
is for later generalizations. These generalizations involve specifying the dynamics of collision process
especially for high energies by giving the analytic properties. Then one relates one experiment to
another using general relations (called dispersion relations) which assume only general properties of
quantum mechanics such as causality, completeness, etc.

Generalization to Compact operators


The single separable potential (in each `) leads to at most one bound state or one resonance. In
general more can be obtained. One way of extending this discussion is to take a large number N of
such terms for each `
N
h̄2 X
N
v` ≡ − gn` (r)gn` (r0 ).
2m n=1

Cases for which || v − v N ||→ 0 as N → 0 are called compact operators. For the above v`N one solves
for eiδ` sin δ` and gets
P
nm gn` (k)Cnm (E)gm (k) N
eiδ` sin δ` = = (+)
2 ∞ dk0 k02 gn (k0 )gm (k0 ) D (k)
R
det[δnm − π 0 k02 −k2 −i ]

Cnm
where det() is the inverse of the N X N matrix 1− < gn | G` | gm > . n Here we see D(+) (k) is a
determinant of an N × N matrix - in the N → ∞ limit it is the Fredholm determinant.
The analytic properties of the generalization is simply obtained by including more bound state
poles and resonances!

n A formal rendition of this structure is, starting from the Lippman-Schwinger equation T = V + V G T, we get
0
(1 − G0 V ) T = V, and T = (1 − G0 V )−1 V. The det(1 − G0 V ), is then the Fredholm determinant.
Chapter 8

Spin

8.1 Introductory Remarks


Our discussion last term involved the notion of a wave function Ψ(~r, t). That idea arose from the need
to formulate a mathematical description of the uncertainty principle, which took the form of Fourier
analysis and construction of wave packets. The wave packets moved according to Schrödinger’s
equation. The appropriate classical limit showed us how to introduce forces and potentials. The
application to bound and continuum states was made and we generated useful stationary states,
which provided a “complete” basis of states. The observation of eigenvalues, and the associated
probability amplitudes, etc. was discussed as a basic “expansion postulate.” Wave packets were
used to provide a rigorous description of scattering.
In all of this, we used Dirac notation as a means for expressing the Fourier transformations in a
neat notation; we expressed operators in a Hilbert space of states and used momentum space and
configuration space representations.
The question arises: is this really a “complete” description - the answer is no! The existence
of internal properties such as the intrinsic magnetic moment and the associated spin leads us to
generalize our discussion.
We shall now discuss the notion of spin - and we shall see that a generalization is needed - we
find that Dirac’s formulation is not simply a shorthand for Fourier transforming, but provides the
basic language for generalizing Quantum Mechanics to handle concepts, such as spin (and later of
isotopic spin and other new quantum ideas, strangeness · · ·)

8.2 Spin
The historical development of the concept of electron spin is a fascinating and illuminating story; it is
told admirably well in M. Jammer’s book “The Conceptual Development of Quantum Mechanics”§3.4.
The concept originated in the observation of fine structure of spectral lines; one observed splittings
of lines in Magnetic (Zeeman effect) and Electric (Stark effect) fields. These effects were recognized
as being “intimately related” to the origin of multiplet structures of spectral lines. Sommerfeld and
later Landé had developed the “vector model” for understanding multiplet structure and thereby
introduced the notion of “space quantization;” i.e. that angular momenta pointed in certain discrete
directions and combined accorded to discrete addition rules. These efforts were aimed at understand-
ing the selection rules for line intensities that were lacking in Bohr’s original theory. To explain the
multiplet splitting (existence of doublets and triplets) the idea was developed that there was an
interaction between the core and valence electrons in alkali atoms responsible for such multiplet
splittings. (That idea is later replaced by the −µ · B = −gµ0 σ · B or σ · ` Thomas spin-orbit splitting
explanation.)
To study these ideas, Stern and Gerlach in 1922 (Z. für Physik 9 349 (1922)) proposed an
experiment to test the Sommerfeld-Lande “magnetic-core” theory of spectral lines and the idea of
space quantization.
CHAPTER 8. SPIN 181

8.2.1 Stern-Gerlach Experiment


This experiment was done using a narrow beam of neutral silver atoms, obtained by evaporation
in a heated oven, with a high vacuum (of 10−4 to 10−5 mm Hg). The beam was passed through a
“DuBois” magnet which provides a strong field gradient at a sharp edge of the pole piece.

S
p p s Deposit 2
p p p p
@
p p p
p p p p
@
p p p p
 -p p p p p p p p p p p p p p p
@
p p p p
@
R I
@
p p p p
p p ps
N Deposit 1
Oven Atomic Beam Magnet
Glass plate

Figure 8.1: The Stern-Gerlach Experiment (SGE) setup

Figure 8.2: SGE Magnet in 3-D

As a result, it was found that a thin deposit formed at only two places, in contrast to what is
expected classically.
The mechanism of interaction with the magnetic field gradient, from a classical viewpoint, is
simply a magnetic moment interaction:

~ µ · B)
~ ∂
F = −∇(~ Fi = −µj Bj (~r).
∂xi
The orientation of the magnetic moment µ~ of the silver atoms with respect to the magnetic field can
take on any value classically and hence one expects a distribution vastly different from the above
result. One expects simply

Uniform Deposit
Figure 8.3: The Stern-Gerlach according to classical theory which predicts a uniform distribution of
Silver on the glass plate as shown above.
CHAPTER 8. SPIN 182
Deposit 1

Deposit 2

Figure 8.4: The Stern-Gerlach Experimental Result. Only two locations have deposits.

The original interpretation of this amazing result was that only certain directions between µ
~ and
~ occurred, which verified the existence of space quantization. Also, it was established from the
B
deflection that the magnetic moment of silver atoms is one Bohr magneton
eh̄
µ0 = = .58 × 10−8 ev/gauss,
2mc
where m is the electron mass. For B ∼ 10+3 gauss, the energy change is about ∆E ∼ µ0 B ∼
.59 × 10−5 ev which is down by ∼ 10−5 compared to Bohr energies ∼ 13.6ev.
(Other experiments were performed later using H, Na, K, Cd, Tl, Zn, Cu and Au).
A puzzle remained, however, in that there were only two deposits. An early explanation for
alkali atoms was to attribute no angular momentum to the valence (s) electron and ± 21 h̄ angular
momentum to the atomic core made up of the remaining Z − 1 electrons plus the nucleus of charge
+eZ. That “explanation” however provided no insight into the question of the mechanism needed
for atoms to be “space quantized” (a point stressed by Einstein and Ehrenfest).
Later developments, which clarified the interpretation of the Stern-Gerlach experiment, are dis-
cussed in Jammer. These involved the realization by W. Pauli of the underlying double-valuedness,
the need for an exclusion principle, and finally the idea of spin, which was invoked only reluctantly.
That step was taken by Uhlenbeck & Goudschmidt (1925).
Another aspect of the Stern-Gerlach experiment, which has been emphasized originally by Von
Neumann (and later by Schwinger ) is its function as a measuring device. We have a special separa-
tion of deposits on the glass and a macroscopic-sized magnet. Using these macroscopic observations,
we are led from a spatial separation to the observation of a microscopic quantum mechanical notion
of magnetic moment and its associated spin. It is in fact always true that “measurements” involve
macroscopic devices and hence involve a realm of classical physics. A careful analysis of the ability
to use classical measuring devices to measure new quantum degrees of freedom is the subject of
Gottfried’s Chapters IV and V. See also Schwinger’s lectures and Sakurai’s book. We shall discuss
this later. For now I should like to point out the need for a wave packet analysis where the incident
beam ( described later as an ensemble using the density matrix) is split into two spatially separated,
non overlapping wave packets

S
p p ~Deposit 2
p p p p
@
p p p p x
p p
@
p p v
q r s p p pp pp pp
p tp p p p p p u
@
p p v
p p p p
p p x
p p p p
p p~
N Deposit 1
Oven Atomic Beam Magnet
Glass plate

Figure 8.5: The Stern-Gerlach setup with wave packet spreading. If the spreading is too large over
the size of the experiment, then the distinct deposits will not appear.

To have no overlap of the wavepackets when they hit the glass, spreading must be small. Small
CHAPTER 8. SPIN 183

spreading is provided here by the mass of the Ag atom. The spreading factor is:
h̄ L h̄ 1 L λ –AG
e ∼ (∆k)2 t ' = ( )( ).
M v M (∆x)2 ∆x ∆x

The DeBroglie wave length for silver


–AG ∼ h̄
λ ,
M v
where M = mass of atom, which  m = mass of electron; ... λ –AG  λ –electrons . The silver atoms
experience much less spreading than for electrons of the same speed.
In a Stern-Gerlach setup, electron wave packets would overlap and interfere ... the spreading
factor eAG  eelectrons . Polarized beam of electrons can not have their polarization measured by a
S-G method. A full analysis of this aspect of the S-G experiment is given in Gottfried; we will not
spend much time on this point.
Of course, one also needs to have a neutral beam to avoid electric field deflections, which is why
electrons, protons, etc. can not be used in a Stern-Gerlach setup.

Magnetic Moment and Spin


The Sommerfeld explanation of fine structure in hydrogen, which he attributed to relativistic cor-
rections to Bohr’s model, was later (1920) applied to the spectra of the alkali metals. These alkali
metals can be thought of as one valence electron moving about a “core” consisting of the other
electrons and the nucleus. Although the observation of fine structure in the level spectrum of these
atoms seems to agree with the Sommerfeld explanation, it was not reasonable, since for these atoms
it could be shown, using classical assumptions of the Bohr model, that the electrons were not very
relativistic. The idea of a magnetic moment for the core could be invoked, but that led to no insight
(Ehrenfest & Einstein’s comments - see Jammer).
Also, in some cases, there arose a need to invoke half-integer quantum numbers to generate the
observed spectrum - that was unusual in view of the quantization procedures of Bohr & Sommerfeld.
The resolution of this dilemma was provided by Uhlenbeck and Goudschmidt (with Ehrenfest’s
encouragement). Before the appearance of Schrödinger papers, Uhlenbeck and Goudschmidt in
1925 suggested that level shifts occur because of an interaction between the “intrinsic spin” (and
its associated magnetic moment) of the electron and the effective magnetic field (B ~ e ) field by the
moving electron. Since Be ' ~`, this leads to a spin-orbit, ~s · ~` interaction. (See later for a better
derivation).
The advantage of their idea was that it has a broad application, not only to alkali atoms but
to lighter & heavier atoms as well. In addition, once accepted by Pauli, the idea provided a firmer
understanding of the exclusion principle, which is so basic to our understanding of the periodic table
and both the atomic and nuclear shell models. The wide applicability of this idea serves to make
it believable - it serves to unify and express concisely a vast array of observations and to permit
predictions. What more can one ask of a reasonably physical concept?
(Note: Pauli (and Lorentz) original objected to the spin notion since a finite sized electron leads
to a violation of relativistic restrictions - here we associate spin with a point particle).
If we accept the notion of spin, we need then only develop a language or formalism to express
the idea of double-valuedness: That formalism consists of inventing an abstract space - spin space
and the spin operators: ~sop = spin represented by a vector Hermitian operator = ~sop† .
We take ~sop to be Hermitian since it is observed indirectly by virtue of its interaction with the
Stern-Gerlach experimental device. The spin operator is understood to act on states in an abstract
Hilbert space.
3 2
~sop2 | s ms >= s(s + 1)h̄2 | s ms >=
h̄ | s ms >,
4
where we now use Dirac notation to formulate the idea of spin-space. Here s is a number, an
eigenvalue, which is s = 12 for the electron. Another eigenvalue equation corresponding to the
~ is
component of ~s along B
1 1
sop
z | s ms >= ms h̄ | s ms >⇒ ± h̄ | s ± >,
2 2
CHAPTER 8. SPIN 184

where ms = + 12 (spin up) or ms = − 12 (spin down) is taken as a means of expressing the possibility
of having only two directions for the spin- corresponding to observing only two deposits of silver in
the Stern-Gerlach experiment 2s + 1 = 2 = even number of m` values = number of deposits. For an
orbital angular momentum, we get 2` + 1 = odd integer only. The nonzero m` = ±1 values should
yield two regions of deflected beam, and the third value m` = ±0 should yield a difficult to isolate
nondeviated beam.
Having made a start on formulating the notion of spin, let me interrupt this direction and make
some comments on the relationship between spin and the associated magnetic moment.

8.2.2 Magnetic Moment


From a classical viewpoint, we can see that a simple relationship exists between angular moment of
a charged particle and its magnetic moment. For an electron we have the classical result that:

~` eh̄
~ = −µ0
µ with µ0 = + = Bohr magneton.
h̄ 2mc
To obtain this result, consider the electron spread over a ring and moving in a circle of area
A = πr2 , and creating a current I, Then the magnetic moment is = Ic A; hence, taking a small part
of the charge dq its contribution is
dq ω 2 dq 1
d~
µ= πr = ( ) (dm) ωr2 ẑ.
c 2π dm 2c
dq e
Since dm = −m , we get
e ~
µ=−
d~ d`,
2mc
and finally
e ~` ~`
~ =−
µ h̄ = −µ0 .
2mc h̄ h̄
Another important way to see the relationships between orbital angular momentum and magnetic
moment is the following. Consider a charged particle moving in a uniform magnetic field. Here the
~ = 1 (B
vector potential A ~ × ~r ), since this yields B~ =∇~ ×A
~ = B.
~ Note that ∇~ ·A
~ = 0. Hence the
2
Hamiltonian is for an electron (q = -e)

1 q~ 2 p2 q ~+A
~ · p~)
H= p − A)
(~ = − p·A
(~
2m c 2m 2mc
e2 1 2 2 2
+ B r sin θ
2mc2 4
The last term (∝ B 2 ) will be needed for the quadratic Zeeman effect and for explaining diamag-
~ = (~
netism (see later). Using p~ · A ~ +A
p · A) ~ · p~ = 2A)
~ · (~
p, we have

p2 e 1 ~
H= + 2 (B × ~r) · p~ + 0(B 2 )
2m 2mc 2

p2 eh̄ ~ ~` p2
= + B · + 0(B 2 ) ∼
= ~ ·µ
−B ~.
2m 2mc h̄ 2m
eh̄ ~ ~
Again we obtain µ ~ = − 2mc ` = −µ0 h̄` for orbital motion in a constant magnetic field. That result
coincides with the previously deduced classical result for orbital motion in a magnetic field.
Now for spin, a similar expression is needed. The above result yields a z - component of angular
momentum and hence of magnetic moment of only certain discrete values.

µz = −µ0 m`

We expect 2`+1 = odd number of “space quantized” directions and hence an odd number of deposits
in the Stern-Gerlach experiment.
CHAPTER 8. SPIN 185

The notion of 12 integer spin leads to only two values if we take the relationships between spin
and angular momentum to be
~s
~ = −µ0 gσ
µ

or
µz = −µ0 gσ ms
yields two values ms = ± 21 .
The “g-factor” gσ for the electron can be measured using either the Zeeman effect or the Stern-
Gerlach measurement of the magnetic moment of Ag as one Bohr magneton. The last method
requires that the magnetic moment of alkali atoms arises from the valence electron only.
To get a µz = ±µ0 for an electron the g-factor for the electron must be gσ ∼
= 2. The corresponding
quantity for orbital motion µz = −g` µ0 ~`/h̄ was seen to be g` = 1 from both a classical and quantum
point of view. The fact that the magnetic moment of the electron is proportional to the spin angular
momentum illustrates the purely quantum mechanical nature of spin. Despite its origin in the
formalism of quantum mechanics, the electron’s spin plays an essential role in our understanding of
the universe.
The g-factor of the electron can be measured very precisely using magnetic resonance techniques.
The value deviates from the above gσ = 2 and the experimental result for gσ is
α
gσ = 2[1 + + · · ·] = 2.00232.

The explanation for this result and a similar situation for the magnetic interactions in atoms (which
is part of the Lamb shift) is a subject dealt with in the quantum electrodynamics developed by
Schwinger, Feynman and Tomonoga.

Magnetic Moment of an Atom


It is of interest to consider the magnetic moment of an atom in a constant magnetic field and examine
the possible role played by the “core” (Z − 1 electrons plus the nucleus of charge +Ze) versus that
of the valence electron.
For example, consider an electron (-e) plus a proton (+e), [or a valence electron plus a “core” of
net charge +e. The magnetic field on each “particle” is determined by

~ 1 ) = 1 (B
A(r ~ × ~r1 ); ∇1 × A(~
~ r1 ) = B1 = B
2

~ r2 ) = 1 (B
A(~ ~ × ~r2 ); ~ 2 × A(~
∇ ~ r2 ) = B2 = B
2
Note: we take B constant over the atomic size ∼ 10−8 cm, which is valid for the inhomogeneous field
set up in the Stern Gerlach gradient. As the atom moves in the inhomogeneous S-G field it feels
a “slowly” varying B-field. The reaction of the atom’s wave packet to a slowly varying B-field is
discussed in Gottfried p 166
The Hamiltonian is thus given as
2 ei 2
X pi −
(~ c A(ri ))
H=
i=1
2mi

p212 P2 e1 ~ ~ e2 ~ ~
H= + − B · `1 − B · `2 + 0(B 2 ).
2µ 2M 2 m1 c 2 m2 c
using ∇i · A(ri ) = 0 we get
p2 P2 ~
H' + −µ
~ ·B
2µ 2M
and
e1 h̄ ~`1 e2 h̄ ~`2 ~`1 ~`2
~ = −[
µ + ] = −µ1 − µ2
2m1 c h̄ 2m2 c h̄ h̄ h̄
CHAPTER 8. SPIN 186

This is the magnetic moment due to the orbital motion of particles 1 and 2 about some arbitrary
center. Consider 1 to be the valence electron and 2 the core. The possibility of an intrinsic magnetic
moment for the electron (or for the “core”) has not been introduced here!
We can introduce the relative and center of mass coordinates
~ + ~r12 m2
~r1 = R
M
and
m1 ~
p~1 = P + p~12
M
r2 = R ~ − ~r12 m1
M
m2 ~
p~2 = P − p~
M
m1 ~ m2 ~ m1 m2
`1 = L+ `+ ~r12 × P~ + R
~ × p~12
M M M2
m2 ~ m1 ~ m1 m2
`2 = L+ `− ~r12 × P~ − R
~ × p~12
M M M2
(e1 + e2 )h̄ L m2 m1 ~`
~ = −µ1 `1 − µ2 `2 = −
µ − (e1 + e2 )
2M c h̄ m1 m2 2M c
+ two additional terms. The additional terms are zero on the average because the expectation values
< ~r12 >= 0 and < p~12 >= 0, by parity.
Also for a neutral atom e1 + e2 = 0 and we get

e1 h̄ m2 m1 ~`
~ =−
µ ( − )
2M C m1 m2 h̄
if e1 = −e of electron, we get

eh̄ mp me ~` ∼ eh̄ ~`
~ −→ −
µ ( − ) =− = −µ0 ~`
2M C me mp h̄ 2me c h̄

since mp  me and M = me + mp ' mp .


Thus the magnetic moment of the atom, assuming no intrinsic electron or core contribution, is
that due to the relative orbital motion of the electron!
Application of this result to the one-valence electron for alkali atoms, we have a 1s orbital,
hence < ~` >= `(` + 1) = 0... no magnetic moment arises from the orbital motion. The magnetic
moment must therefore arise either from some magnetic moment of the core or from an intrinsic
magnetic moment of the electron. The magnetic moment of the core for sodium would have to be
double-valued in µz , which would not P arise from combining the orbital motion of the constituent
core
electrons of the core. This is, Lcore = i=1 `i would still yield 2 L+1 = odd. It is therefore hard
to believe that the core would explain the double-valued µz seen in the Stern-Gerlach experiment
unless some intrinsic magnetic moment were assigned to the electrons. A simple procedure is to
assign an intrinsic magnetic moment to the valence electrons. (Also, it can be shown that closed
shells give as much +µz as −µz and hence have zero magnetic moment - as is observed for the inert
gases.)
The above discussion shows that the electron’s orbital motion yields a magnetic moment, but
to yield the correct double-valued magnetic moment µ ~ · ẑ for alkali metals one needs to assign an
intrinsic moment to the electron. The idea of assigning a core magnetic moment of double value is
rejected as discussed in the preceding paragraph on physical grounds. Also an assignment of a core
magnetic moment is an “ad hoc” assumption that could perhaps account for this result, but lacks
the universal applicability of the notion of an intrinsic electron magnetic moment.
CHAPTER 8. SPIN 187

8.2.3 Spin of an Electron


Once it is established that the electron has an intrinsic magnetic moment, we can now relate that
property to the intrinsic angular moment called the spin. (The spin of a particle or of a system such
as a nucleus can be defined as its angular moment as determined in its rest (P~ = 0) frame - this
definition holds for relativistic field theory cases except for massless particles.)
The step of attributing an intrinsic spin to a point electron is nontrivial. (Pauli originally rejected
that notion since for a spinning system with nonzero angular momentum, a size was involved that
led to speeds at the electron’s edge greater than c.) That difficulty is circumvented by assigning a
spin to a point electron using the relationship

~s
~ 0 = −gσ µ0
µ ,

which has been verified by other experiments. a
The spin operator has definite properties that we assign to it to express the basic idea of a
double valued or spin= 12 system subject to the laws of quantum mechanics. Thus we develop the
language or formalism to meet a need to express and use the concept suggested by Uhlenbeck and
Goudschmidt.
Our first observation is that we want ~s to be an observable corresponding to an angular momen-
tum. Hence ~s is a vector and Hermitian operator

~s† = ~s ;

where there are really three operators sx , sy , sz . Here we understand these operator to operator in
an abstract Hilbert space. There is also the fundamental property of quantum mechanics which we
assign to these operators; namely, the uncertainty principle in the form of the commutator

[sx , sy ] = ih̄sz , etc.

This form was found earlier to hold for orbital angular momentum as a consequence of ∆px ∆x ≥ h̄/2;
now we introduce it for spin operators as an independent statement. The consequence of this
expression of our inability to measure all aspects of ~s simultaneously is to give meaning to the vector
model and the space quantization of Sommerfeld and Lande. We can understand discreteness in
spatial orientation as a result of the limitations placed by h̄ on our ability to measure small action
systems.
With ~s 2 ≡ s2x + s2z + s2y , we get
[ s2 , si ] = 0.
Along with the above cyclic commutators, one finds only s2 and sz can be simultaneously measured.
Determination of sz disturbs the measurement of sx and sy , but does not alter the measurement of
a precise ~s2 .
Since only ~s2 and sz are really simultaneously observable, we can label our spin states by the two
associated quantum numbers. A variety of notations are used - let me provide a typical list used to
represent the spin-states.
1 1
| s ms > or | > | − >
2 2

χsms (1)

1
α(1) =| >=↑1 spin − up
2
1
>=↓1 spin − down
β(1) =|
2
See later for the matrix representation of the spin operators and the spin states. Here the label 1
is used to indicate these are the spin states of particle 1. Another particle, particle 2, would be
labelled as α(2), β(2), etc.
a From now on the spin operator ~
sop will be written simply as ~s, where the operator aspect will be clear from its
usage.
CHAPTER 8. SPIN 188

The uncertainty principle commutators can also be expressed in terms of raising and lower
operators with
s± = sx ± isy
and
[s± , sz ] = ∓s± h̄.
This last commutator has the consequence that

< s m0s | s± sz − sz s± | s ms >= ∓h̄ < sm0s | s± | sms >

(ms − m0s ± 1)h̄ < sm0s | s± | sms >= 0


Hence s± have nonzero effect only if m0s = ms ± 1. Thus we identify that s± are indeed spin raising
and lowering operators. To make this property full we have the property
r
0 3
< sms | s± | sms >= + − ms (ms ± 1) h̄ δm0s ,ms ±1 .
4
2
This property follows from the definition of ~s and s± :

s2 = s2x + s2y + s2z = (sx + isy )(sx − isy ) + s2z + i[sx , sy ]

... s2 − sz (sz − h̄) = s+ s−


< sm0s | s+ s− | sm0s >=|< sm0s | s+ | sms >|2 δm0s ,ms +1
3
= h̄2 [ − m0s (m0s − 1)]
4
3
= h̄2 [ − ms (ms + 1)],
4
which yields the result above.
For s = 21 , this construction simply expresses the double-valuedness and the raising and lowering
property - i.e.
1 1
s+ | >= 0; s− | − >= 0
2 2
and the spin-flip properties
1 1 1 1
s+ | − >= h̄ | + >; s− | >= h̄ | − > .
2 2 2 2
We are dealing with an object that has two states.
Properties of interest that can either be deduced or simply postulated are:

s+ s+ = s− s− = 0

double valuedness.
Other properties are:
s2 = s+ s− + sz (sz − h̄)
= s− s+ + sz (sz + h̄).
Note that from s2± = 0, we get

s2± = (sx ± isy )2 = s2x − s2y ± i(sx sy + sy sx ) = 0

or
s2x = s2y
and
sx sy = −sy sx
These two results are therefore equivalent to saying s2± = 0 and find that that s2x = s2y and sx sy +
sy sx = 0 express the double-valued property of spin!
CHAPTER 8. SPIN 189

Summary
~s is a vector Hermitian operator subject to the uncertainty principle: [sx , sy ] = ih̄sz ; ... s2 , sz label
states.
Also: [s± , sz ] = ∓s± h̄; s± are raising, lowering operators for spin.
Double-valued property of spin s2± = 0 leads to s2x = s2y and sx sy + sy sx = 0.

8.3 Pauli-Spin Operators


It is convenient to introduce the Pauli-spin operators and to separate out the h̄/2 factor by defining

~s ≡ ( )~σ
2
for each of the three operators σx , σy , σz which are called the Pauli spin operators.
Here ~σ is a Hermitian, vector operator defined to act in the previously defined spin states |
s ms > . Properties of ~σ have already been defined as
σi† = σi Hermitian
[σx , σy ] = 2iσz or ~σ × ~σ = 2i~σ
[σν , σλ ] = 2iνλµ σµ
Uncertainty Principle

σ± = σx ± iσy
Raising and Lowering Operators
1
σ± | ± >= 0
2
Double value
1 1
σ± | ∓ >= 2 | ± >
2 2
Spin flip

σ 2 | sms >= 3 | sms > or σ 2 = 3


σz | sms >= 2ms | sms > or σz2 = 1
Eigenvalues
unimodular
σx2 = σy2
σ 2 = σx2 + σy2 + σz2 = 2σx2 + 1 = 3 ... σx2 = 1
... σ 2 = σ 2 = σ 2 = 1
x y z

σx σy = −σy σx = iσz
generally
σµ σλ + σλ σµ = 2δλ,µ σu µσλ = iµλν σν uµ 6= λ
σx σy σz = i
The trace in spin space is also of interest,
X
T r ~σ ≡ < s ms | ~σ | ms >= 0
ms

Finally, it is easy to show that (see H.W.)


~ σ · B)
(~σ · A)(~ ~ =A
~·B
~ + i~σ · (A × B)

~ and B
where ~σ refers to the spin of a single electron and A ~ are vectors, which could be operators
that commute with ~σ .
CHAPTER 8. SPIN 190

8.3.1 Pauli-Spin Matrices


These operator properties can also be represented by the Pauli-Spin matrices, where we identify the
matrix elements by  
0 1 0
< s ms | σz | s ms >−→ (8.1)
0 −1
Note that
σz = σz†
Similarly for the x− and y−component spin operators
 
0 0 1
< s ms | σx | s ms >−→ (8.2)
1 0

σx = σx†
and  
0 −i
< sm0s | σy | sms >−→ (8.3)
i 0
with Hermiticity again
σy = σy†
Also  
0 2
σ+ = (8.4)
0 0
 
0 0
σ− = (8.5)
2 0
Note that these 2 x 2 matrices are traceless Tr[~σ ] = 0, unimodular σi2 = 1 and have unit
determinant | det σi |= 1. Along with the unit operator they form a complete set of 2 x 2 matrices
and hence any 2 x 2 matrix can be expressed in terms of the unit matrix

 
1 0
1≡ (8.6)
0 1

and ~σ . Later, we shall introduce a density matrix ρ = a +~b · ~σ = a + b~n · ~σ to describe an ensemble of
particle spin directions as occurs in a beam of spin-1/2 particles. Also a spin dependent amplitude
for a spin-1/2 on a spin-0 scattering process will be introduced in the form

f (~k 0~k) = g 1 + hn̂ · ~σ .

8.4 Complete Orthonormal Spin-States


We have already tacitly assumed that the spin states | s ms > are normalized and orthogonal, but
let me make these statements explicit. The normalization and orthogonality are given by

< s ms | s m0s >= δm0s ,ms .

Normalization just means that the probability of finding a spin-up electron with spin-up is unity.
The orthogonality means that the spin-up and spin-down situations are distinct physical situations,
which do not interfere and can be separated; this property follows from the existence of an eigenvalue
equation which asserts the existence of a physical attribute.

sz | s ms >= h̄ms | s ms >,

which can be precisely measured. To show the connection between the eigenvalue equation and
orthogonality, consider

< sm0s | (sz − sz ) | s ms >= 0 = h̄(m0s − ms ) < s m0s | s ms >


CHAPTER 8. SPIN 191

< s m0s | s ms >= 0 if ms 6= m0s .


Now we can introduce the idea of completeness for these spin states. We must recognize that
completeness is a physical concept for which a mathematical language is developed. Earlier, we
discussed completeness for the wave function Ψ(~r, t) and could even deduce it from the Schrödinger
equation; however, with the omission of spin degrees of freedom that “completeness” was incomplete.
We need to add in the spin degrees of freedom!
The spin of an electron can point up or down. That does not exhaust the possibilities since a
general spin-state can be expressed as a linear combination or superposition ( called a qubit) of the
spin-up and spin-down situations
1 1 1 1 X
| χa >= a+ 12 | , + > +a− 21 | , − >= ams | s ms > .
2 2 2 2
ms =± 12

Here we used the quantum mechanical notion of superposition. The coefficients | a1/2 |2 and | a−1/2 |2
give the probability of finding the spin-up or down in the general spin state | χa >. The electron
has only a total spin of s = 1/2, no spin excited states exist; therefore, no alteration in s-value is
incorporated into the formalism here.
Normalization requires that | a+ 21 |2 + | a− 21 |2 = 1 or
X X 1 1
| ams |2 = 1 = < χa | ms >< ms | χa >=< χa | χa >
ms ms
2 2

1
ms | xa >= ams
<
2
1
< xa | ms >= a∗ms .
2
Thus we see our completeness can be expressed in the closure form
X
| sms >< sms |= 1
ms

When we later discuss the rotation group, we will see that


X
| χa >= ams | s ms >
ms

an eigenstate of ~σ · n̂, where n̂ is the direction in which the spin points after a rotation from the
original axis. This rotation os the spin state will play an essential role in describing an ensemble of
particles with various spin directions.
The coefficients can also depend on time ams (t), which corresponds to changes in the electrons
spin such as occurs in precession. A corresponding Schrodinger equation is then used: H | xχa >=

ih̄ ∂t | χa > . This is equivalent to
X ∂
< ms | H | m0s > am0s (t) = ih̄ am (t),
∂t s
m0s =± 12

which represent two coupled equations for the a+ 21 and a− 12 coefficients. We shall return to this
problem later.

8.4.1 Separable and Nonseparable Space-spin States


The full description of the electron now requires specifying the probability amplitude for finding an
electron at a given point in space ~r and time t, as well as giving the probability amplitude for finding
it with spin-up or spin down (a± 21 ). Hence, the wave function is extended, along with the associated
normalization, orthogonality and completeness properties, to include spin. This is accomplished by
including the spin-space or spin state as:

Ψαsms (~r1 ζ1 , t) = Ψα (~r1 , t)χsms (ζ1 ).


CHAPTER 8. SPIN 192

Here α labels the wave function, such as with the ~k0 of the wave packet centroid, or for bound
orbitals n` quantum numbers. The spin “coordinate” ζ1 is simply a label to remind us that this is
the spin for particle 1. The spin state, which is described in several equivalent notations: χsms (ζ1 ) ≡|
sms >1 ≡< ζ1 || sms >, is simply tacked on to the spatial wave packet Ψα (~r1 , t). Therefore this is
a separable form, which is our stating point. The general space-spin state, wherein we stipulated
the systems location in time along with its spin aspects, is not in general separable. For example
if we impinge a separable spin state on to a Stern-Gerlach measuring device, we wind up with
two separated wave packet with different spin direction, i.e. the final state is of the form ψ+ (~r) |
upspin > +ψ− (~r) | downspin >, where the two wave packets are localized at the distinct regions
of silver deposition. Thus in the SGE the final wave function is not separable, but is a sum of two
separable terms. This generation of a nonseparable space-spin state from an initial separable is also
characteristic of the scattering process, indeed of all interactions. we can say the the spatial and
spin aspects get “entangled.”
Let us return to a case where the spin state is in the pure spin state χs ms . The above separable
wave function is described by:

Ψαs ms (~r1 , ζ1 , t) = Ψα (~r1 , t) χs ms (ζ1 ) = Ψα (~r1 , t) < ζ1 | s ms >=< ~r1 , ζ1 | Ψα (t)s ms > .

For a general spin state, χa , which is a linear combination of ↑, and ↓ spin states, we have
Ψα a (~r1 , ζ1 , t) = Ψα (~r1 t)χa (ζ1 ), which is still separable but with a general orientation of the spin
labelled by a.
The momentum space version of the above separable wave function is

Φαs ms (~
p1 ζ1 t) = Φα (~
p1 t)χs ms (ζ1 )

or
p1 t) < ζ1 | sms > .
= Φα (~
A representation independent expression, which leads to the above wave function in either the ~r
or p~ representations, is
| Ψαs ms >=| αs ms >=| Ψα >| sms > .
For a product or separable spin state, we write:

| Ψα a >=| Ψα >| χa > .

The important point is that spin can be tacked-on by taking a “product” of two Hilbert space vectors
each in a separate space | Ψα > in the configuration space of r3N dimensions times the spin state
| Xχa > in the 2-dimensional spin space.
The normalization, and orthogonality properties are now generalized to include the spin:

< α0 s m0s | αs ms >= δα α0 δm0s ms

< Ψα0 | Ψα >< sm0s | sms >= δαα0 δm0s ms


Z
= Ψ∗α0 (~r1 t) Ψα (~r1 t)d~r1 < sm0s | sms >
Z
= Ψ∗α0 sm0 0 (~r1 ζ1 t) Ψα s ms (~r1 ζ1 t) dτ1 ,
s

where we now assume that the spatial states are orthonormal (i.e. not wave packets). The “integra-
tion” symbol dτ1 includes d~r1 over the continuum, and a dot product over spin states - the label ζ1
is an artifice to remind us to take norm of the vector in the full Hilbert space.
The completeness statement for expanding a general nonseparable state | Ψ > in terms of the
separable spin 1/2 particle wave functions | Ψαsms > (non-relativistic) is:
X
Ψ(~r1 ζ1 t) = < α s ms | Ψ > Ψα (~r1 t) χs ms (ζ1 ).
α ms

This leads to the closure form


X X
| Ψαsms >< Ψαsms |= | Ψα >| sms >< sms |< Ψα |= 1.
αms αms
CHAPTER 8. SPIN 193

Note the state of an electron is described in general by a nonseparable wave function Ψ(r1 ζ1 t),
which has been expanded in terms of the complete (separable) wave functions | Ψα >| sms > . The
spin and spatial degrees of freedom are usually coupled, for example by an interaction

−µ0~σ · B(r1 ).

As a result of such coupling, an initial separable state | Ψα >| sms > typically evolves in time
in both its spacial part | ψα > and in its spin part | sms > . Motion through an inhomogeneous
field would then be described by a deviated wave packet and a precessing spin. See later for another
example of coupling of various | ψα >| sms > states by scattering.
CHAPTER 8. SPIN 194

8.4.2 Two Spin-1/2 Particles


Let us now consider the case of two spin-1/2 particles (two electrons or two protons, say). With
each particle we associate Pauli spin operators and, of course their associated spin
h̄ h̄
~s (1) = ~σ (1) ≡ ~σ1
2 2
refers to particle 1, and
h̄ h̄
~s (2) = ~σ (2) ≡ ~σ2
2 2
refers to particle 2.
Here ~σ (1) and ~σ (2) are operators in the Hilbert space associated with the spin states | sms >1
and | sms >2 of particles 1 and 2, respectively. Now each spin satisfies the requirements of being a
vector, Hermitian operator, subject to the uncertainty principle and a double-valuedness, and have
2ms = ±1 only. Hence, for each electron the Pauli operators are recovered. The two spins are, in
addition, both simultaneously observable since they are fixed at a value of 12 h̄. That is the two spins
are distinct, observables and there operators commute

[~s (1), ~s (2)] = 0

which states that the spins are mutually compatible, i.e. distinct (separate) physical attributes.
Here we ignore the unrealized possibility of one electron disturbing another in such a way as to
“excite” it to a higher spin. b
Since the two spins are separable, we first consider the four product states
1 1
|s + >1 | s + >2 ≡ α(1)α(2)
2 2
1 1
|s + >1 | s − >2 ≡ α(1)β(2)
2 2
1 1
|s − >1 | s + >2 ≡ β(1)α(2)
2 2
1 1
|s − >1 | s − >2 ≡ β(1)β(2), (8.7)
2 2
where we have reverted to the notation alpha(i) denotes spin up and beta > (i) denotes spin down
for the ith electron.
The total spin of the two electron system is defined by
~ = ~s (1) + ~s (2),
S

from which it follows that

Sz α1 α2 = 1 h̄α1 α2 (8.8)
Sz α1 β2 = 0 h̄α1 β2
Sz β1 α2 = 0 h̄β1 α2
Sz β1 β2 = −1 h̄β1 β2 .

These satisfy the eigenvalue equation for the z-component of the total spin

Sz | SMS >= h̄MS | SMS > .

We can also ask for combinations of the above product states that have the additional property of
being eigenstates of the total spin angular momentum defined by

~ = ~s(1) + ~s(2) = h̄ (~σ (1) + ~σ (2))


S
2
b That is reasonable to ignore since it never occurs. It is not so ridiculous, for example for the two nucleon system

there are excited spin states, called nucleon isobars, and they can excite each other. For example, the 32 − 32 resonance
is a nucleon with “spin” 23 .
CHAPTER 8. SPIN 195

~ = 0. Hence we can define the spin-states of the


It is easy to show that [Sx , Sy ] = ih̄Sz and [S 2 , S]
two spin 1/2 particles as simultaneous eigenstates of S 2 and Sz
2
Sop | SMS >= h̄2 S(S + 1) | SMS >

Sz | SMS >= h̄MS | SMS >,


where S(S + 1) and MS are the eigenvalues.
To construct these eigenstates note that
3 2
S2 = h̄ + 2~s(1) · ~s(2)
2
We can immediately arrange the product spin states into three classes with respect to the MS =
ms1 + ms2 eigenvalue:
MS = 0 α(1)β(2) and β(1)α(2)
MS = +1 α(1)α(2)
MS = −1 β(1)β(2)
This is equivalent to

Sz | sss >1 | sms >2 = h̄(ms1 + ms2 ) | sms >1 | sms >2

... MS = ms1 + ms2 .


Using the fact that
1
~s(1) · ~s(2) = sz (1)sz (2) + [sx (1) + isy (1)][sx (2) − isy (2)]
2
1 1
= sz (1)sz (2) + [s+ (1)s( 2) + s+ (2)s( 1)] + [sx (1) − isy (1)][sx (2) + isy (2)],
2 2
let us now construct the spin eigenvalue equation for the total spin (squared)

S 2 | SMS >= h̄2 S(S + 1) | SMS > .

To find these eigenstates | SMS > of both S 2 and Sz , we see that the total spin squared, S 2 =
s21 + s22 + 2~s1 · ~s2 , satisfies

~ = S 2 = 3 h̄2 + 2sz (1)sz (2) + s+ (1)s− (2) + s− (1)s+ (2).


~ ·S
S
2
Note that ~σ · ~σ = 3, for each electron and that the raising and lowing operators also commute for
distinct electrons [s+ (1), s± (2)] = 0. Now consider the total spin squared acting on the four product
states taking into account the raising and lowering properties:
3 1
S 2 α(1)α(2) = [ + ]h̄2 α(1)α(2) = 2h̄2 α(1)α(2) = 1(1 + 1)h̄2 α(1)α(2)
2 2
Hence | S = 1, MS = 1 >≡ α(1)α(2). Similarly, | S = 1, MS = −1 >≡ β(1)β(2).
For the α(1)β(2) and β(1)α(2) product space the action of S 2 is
3
S 2 α(1)β(2) = { h̄2 + sz (1)sz (2) + s+ (1)s− (2) + s− (1)s+ (2)} α(1)β(2) (8.9)
2
3 1
= { h̄2 + 2h̄2 (− )} α(1)β(2) + h̄2 β(1)α(2)
2 4
= h̄2 α(1)β(2) + h̄2 β(1)α(2)

and
3
S 2 β(1)α(2) = { h̄2 + sz (1)sz (2) + s+ (1)s− (2) + s− (1)s+ (2)} β(1)α(2) (8.10)
2
3 1
= { h̄2 + 2h̄2 (− )} β(1)α(2) + h̄2 α(1)β(2)
2 4
= h̄2 β(1)α(2) + h̄2 α(1)β(2).
CHAPTER 8. SPIN 196

MS S=1 S=0

1 α1 α2 -

α1 β2√
+α2 β1 −α2 β1
α1 β2√
0 2 2

-1 β1 β 2 -

Table 8.1: The Spin States for Two Electrons

Combining these two results we get:


α(1)β(2) ± α(2)β(1) α1 β2 ± β1 α2
S2 √ = h̄2 √
2 2
β(2)α(1) β(1)α(2)
± √ + √
2 2
α(1)β(2) ± α(2)β(1)
= h̄2 (1 ± 1)[ √ ]
2
α(1)β(2) ± α(2)β(1)
= h̄2 S(S + 1)[ √ ]
2
We arrive at the singlet (S = 0) and triplet (S = 0) spin states
α(1)β(2) − α(2)β(1)
| S = 0, MS = 0 >= √
2
α(1)β(2) + α(2)β(1)
| S = 1, MS = 0 >= √ .
2
We have then one | S = 0, MS = 0 > and the three triplet states | S = 1, MS = 0, ±1 >, which
are the four possible spin states for two spin 1/2 electrons.

Other properties of | SMS > States


Complete Orthonormal Basis < S 0 MS0 | S MS >= δSS 0 δMS MS0 Completeness
X
Ψ(~r1~r2 ζ1 ζ2 , t) = ψα (~r1~r2 t) | S MS >12 GαS MS
αS MS

plus orthonormal leads to closure statement:


X
| S MS >< S MS |= 1
S MS

Projection Operators - Spin


see HW

8.4.3 Exchange Operators - Spin


see HW
Footnote: Again the general wave function ψ(~r1~r2 ζ1 ζ2 t) is not separable but can be expanded as
on the RHS. As a coupling acts the mixture (CαSMS ) changes depending on the initial conditions
and the form of the coupling term.
CHAPTER 8. SPIN 197

8.4.4 Scattering With Spin


An application of these properties of spin will now be made to the special case of scattering of a
spin 1/2 particle off a spin zero absorbing target. This provides a way to generate polarized nucleon
beams and also illustrated how a separable space-spin initial state evolves into a nonseparable state.
It also show why we refer to the “S-matrix,” since it is a matrix in spin space.
The existence of spin implies a “Polarization of ψ waves” which can be used to define spin in
terms of a double scattering. I will return to that point after describing how to include spin in our
description of scattering.
Consider a spin- 12 particle incident on a spin-less target (represented by a potential). For example,
this situation is realized in nucleon-He4 scattering or electron-atom scattering. If the Hamiltonian
includes a coupling between the spatial and spin-degrees of freedom in the form of a spin-orbit
coupling, we have
p2
H= + VW (r) + (~σ · ~`)VS0 (r).
2m
Such a spin-orbit force can lead to a polarized beam of scattered particles by means of the following
situation:

X s s

 Spin up
s
P
P 
P 
P 

 P
P
~ P
P  

P
 P P


 P
P
P
 X
X s
 PP

 X P
P
P Spin down
P

Figure 8.6: The scattering of a spin 1/2 particle from an absorbing potential (of spin zero). The dot
denotes spin up, the × spin down (into page). The spin orbit potential is assumes to be attractive
when the spin and orbital vectors are parallel.

The sign of ` changes with the impact parameter and hence the up beam and down beam each
have polarized electrons, or at least a high degree of polarization. (Careful analysis shows a strong
absorption is needed to make the above rough description valid). In fact, pHe4 leads to a beam of
nucleons polarized to ≤ 95%. Experiments with polarized beams and/or polarized targets forms a
rich subject in nuclear physics. It is hard to get a beam of polarized electrons, one can do it however
starting from a low energy polarized laser beam that scatters off a high energy unpolarized electron.
The recoil electron maintains high energy and picks up a high degree of polarization. This has been
done at high energy synchrotron facilities.
Let us now analyze how the scattering amplitude (or amplitude of the scattered wave) depends on
angles when spin is included. Our previous analysis of scattering applies, except we must “tack-on”
the spin. A plane wave is now described by
~
eik·~r
< ~r | ~ksms >≡ | sms >,
(2π)3/2
where the spin is now also included.
The Schrödinger equation, including a spin-orbit coupling, can be solved for the stationary state
scattering problem:
[H0 + VW + (~σ · ~`)VS0 ] ψ = E ψ
(E − H0 ) ψ = (VW + ~σ · ~` VS0 ) ψ
eik·r
Z
(+) (+)
ψ~ (~r) = | sms > + (~r | G(k 2 + i) | ~r 0 )U (r0 )d~r 0 ψ~ (~r 0 ),
ksms (2π)3/2 ksms

2m
where U = h̄2
V and
0
1 eik|~r−~r |
(~r | G(k 2 + i) | ~r 0 ) = − .
4π | ~r − ~r 0 |
CHAPTER 8. SPIN 198

| sm0s >< sm0s |


P
The asymptotic wave function is, after introducing 1 = m s0

1 (2π)3 X 0 eikr ~ (+)


Ψ(+)~ksms → {e ikz
| sms > − | s m s > < k s, m0s | U | Ψ~ >}.
(2π)3/2 4π
m0
r k s ms
s

We can pick out the amplitude of the outgoing spherical wave as


2m ~ 0 0 (+)
< sm0s | f (~k 0 , ~k) | sms >= −2π 2 < k sms | V | ψ~ sms > .
h̄2 k

Here < sm0s | f | sms > is the amplitude of the outgoing spherical wave with spin m0s . The
amplitude f is seen to be an operator in spin-space of the form

f = g(~k 0 , k)1 + (~σ · n̂)h(~k 0 , k),

which is obvious in Born approximation, but holds in general for a hermitian potential V. We have
~0 ~
replace the vector ~` by a unit vector |~kk0 ×k
×~
k|
= n̂ which is in the direction of ~`, i.e. perpendicular to
the scattering plane. This is the only scalar that can be constructed from the vectors in the problem.
The cross section that follows from the above amplitude is of the form

σ(θ, φ) = a(θ) + b(θ) cos φ.

The basic point is that the spin 1/2 particles yields a dependence on cos φ, which is the angle between
the spin polarization and the normal to the scattering plane. Other spin values, like spin one or spin
3/2, give a different dependence on cos φ.
We can use the above result for σ(θ, φ) as an operational definition of spin. A spin-1/2 particle
is then determined if it is polarized, then analyzed and varies as cos φ.

8.4.5 Additional Comments about SGE


The Stern-Gerlach experiment (SGE) is of great significance. It represents the basic measurement
process whereby an apparatus, which is of course a classical object, reveals features of the quantum
world. It has been used by J. Schwinger and subsequently by others as the basis for formulating the
symbolic language of quantum mechanics. It is worthwhile looking at his lectures as published in “
Quantum Mechanics-Symbolism of Atomic Measurements.” The title itself tell you that Schwinger’s
approach is based on measurement, i.e. on a brilliant analysis of the SGE.
As an illustration of the significance of the SGE, let us consider the following situation.

S p s Deposit 2
p
p p p p
@
p p p p p
@
p p p
p p p p p pp ppp ppp p p p p p p p p p p p p p
p p p p p ŷp Axis p
@
p p p p
p p p p
p p p p
p p s Deposit 1

x̂ Axis N
Oven Atomic Beam Magnet
Glass plate

Figure 8.7: The SGE with the atomic spin of 1/2 pointing in the x̂ direction. The classical and
quantum predictions are dramatically different.

Consider an incident spin 1/2 system ( such as a Ag atom whose valence electron provides the
magnetic moment) which impinges on a SG setup with the magnetic moment in the same direction
CHAPTER 8. SPIN 199
S p s Deposit 2
p p p
p p p p
@
@
p p p
p p p
p p p p p p pp ppp ppp p p p p p p p p p p p p p
- p p ŷp Axis p
@
p p p p
p p p p
p p p p
p p s Deposit 1
N
Oven Atomic Beam Magnet
Glass plate

Figure 8.8: The SGE with the atomic spin of 1/2 pointing in the ŷ direction. The classical and
quantum predictions are dramatically different.

as the incident velocity, which we take as the +ŷ direction. The magnetic field is in the ẑ direction,
with a gradient also in that direction. Classically, we look at the force
F~ = −∇(−~
~ µ · B)
~ = µz ∇B
~ z = 0,

where the last step follows from the statement that the spin, and hence the magnetic moment, is
definitely in the ŷ direction and therefore is definitely not in the ẑ direction. Such definite remarks
are characteristic of classical physics.
In the quantum statement, we can always pick the ŷ direction as our axis of quantization, or the
direction in which we definitely know there is a spin component of 1/2. Usually it is the ẑ direction
that is so adopted as the direction in which we definitely know there is a spin component of 1/2.
What does this mean? Recall that the commutator [Jz , Jx ] = ih̄Jy provides an uncertainty relation
∆Jz ∆Jx ≥| Jy | h̄2 or ∆Jz ∆φ ≥ h̄2 , where ∆φ ≡ ∆Jx / | Jy | . This uncertainty relation states that
even when we know that the spin is in a state | s, ms = 1/2 > with definite component of 1/2 in the
ẑ direction, as expressed as an eigenvalue equation
σz | s, ms = 1/2 >= +1 | s, ms = 1/2 >,
the x̂ and ŷ component are not determined and the net angle in which the full spin vector points
is completely undetermined. That is in the ∆Jz = 0 limit ∆φ = ∞. That situation is described
pictorially as the “spin-cone” shown below.

z-axis

y-axis

Figure 8.9: A definite value in the ẑ along with uncertainty in the angular momentum in the ŷ and
x̂ directions. A similar picture holds for the cases when there is a definite value in the ŷ direction
along with uncertainty in the angular momentum in the ẑ and x̂ directions.

For the spin definitely in the ŷ direction, the eigenvalue equation is


σy | +ŷ >= +1 | +ŷ >,
where | +ŷ > describes a spin state with ∆Jy = 0. Now it is the Jz and Jx directions that are
uncertain and the “spin-cone” is now pointed in the ŷ direction. Does this situation yield any
magnetic moment effects in the ẑ direction of the magnetic field and therefore to a deflection?
Classically the answer is no, but now the answer is yes. This QM answer is due to the above
uncertainty principle description and to the equivalent statement that the spin state | +ŷ > is a
simple superposition. c The superposition for this case is
1
| +ŷ >= √ (| s ms = 1/2 > +i | s ms = −1/2 >),
2
c Later we will obtain such superpositions for any direction n̂ by Wigner rotation operators.
CHAPTER 8. SPIN 200

which using the matrix rendition


 
1
| s ms = +1/2 > = | +ẑ >→ ,
0

and  
0
| sms = −1/2 > = | −ẑ >→
1
in matrix form is  
1 1
| +ŷ >→ √ .
2 i
Let us check this superposition is an eigenstate of σy
     
0 −i 1 1 1 1
σy | +ŷ >→ √ = +1 √
i 0 2 i 2 i

Matrix multiplication shows the mixture is correct. d


We can now ask: How likely is it that we find the spin pointing up when we know definitely
that there is a spin component of +1/2 in the ŷ direction? That answer is provided by the overlap
|< s ms = +1/2 | ŷ >|2 = 1/2, a 50 % chance. For spin down there is also |< s ms = −1/2 | ŷ >|2 =
1/2, a 50 % chance. Thus for this case the beam is split into two, in striking contrast to the classical
result.
Thus the SGE result reveals the underlying need for the uncertainty principle and the associated
superposition of states. We will consider the SGE with general spin directions and spin greater than
1/2, after we discuss the rotation group.

8.5 Basics
The introduction of spin has led us to use a “Hilbert space” language to describe this quantum-
mechanical attribute. Four properties were built into the language, namely, that the spin is a
vector operator (to make it an angular momentum), that spin is a Hermitian operator (to make
it an observable), that spin is subject to an uncertainty principle (and hence is represented by
noncommutative operators) and finally that spin is double-valued for electrons. From these four
properties the Pauli- operators and Pauli-matrices were constructed.
This language of operators, states, probabilities when applied to spin was presented in a rather
casual fashion. Many of the steps were justified by the wave mechanics discussions of last term.
Then the ideas of Hermitian operators, completeness, orthonormality, probability amplitude were
developed in an explicit fashion using differential equations, functions and differential operators.
With the advent of spin a more general language and notation is required, which is provided by
Dirac’s transformation theory.
There are several ways to develop the formalism needed to incorporate notions such as spin into
quantum theory. We have used a combination of the historical approach and the description of
quantum mechanics from the wave mechanics viewpoint. (In fact, matrix mechanics was developed
first-so our historical version is not altogether accurate.) Another approach is simply to present the
basic postulates, to define the rules of the game, and to explore the physical consequences - such
as predicting what occurs in measurements. A third approach is to recognize from the beginning
that quantum mechanics is based on the measurements of small action systems, which are easily and
uncontrollably disturbed by the act of measurement. Then this idea has to be expressed formally in
terms of a measurement symbol algebra. This last approach is most elegant and the one discussed
by Gottfried in his Chapter 5 and in Sakurai’s book. In fact, it was originated by J. Schwinger, who
d The eigenstates for ±ŷ and ±x̂ cases are
1
 
1
| ±ŷ >→ √ .
2 ±i

1
 
1
| ±x̂ >→ √ .
2 ±1
CHAPTER 8. SPIN 201

develops his approach to basic quantum mechanics in his lectures “ Quantum Mechanics-Symbolism
of Atomic measurements,” which I recommend if you are interested in this measurement algebra
approach. He starts from the SGE as a classical measuring device that reveals quantum laws. I will
make a few comments about this approach later.
I will now simply state the postulates of quantum mechanics and follow Dirac in introducing the
idea of Hilbert space to describe the formalism.
I suggest you read the first few chapters of Dirac - He gives an elegant, physically motivated
rendition of the Hilbert space notation. You might also study Chapter 6 of Schiff, especially the
sections dealing with Dirac’s approach. The basic Postulates deal with:
1. States

2. Observables
3. Eigenstates
4. Expansion Postulate - Basic (CON)
5. Measurements and Probability

6. Time Develop - Dynamics

8.6 Postulates of Quantum Theory


[adopted from Bjorken and Drell]

1. States
The “state” of a given physical system is described by a vector | Ψa > in a complex, linear
vector space.
From | Ψa > all measurable properties needed to describe the state of the system can be
extracted. The properties of this linear vector space will be discussed later.
2. Observables
Every physical observable is represented by a linear Hermitian operator in the complex linear
vector space.
A complete set of commuting Hermitian operators determine a set of observables that can be
simultaneously measured with precision- “mutually compatible” observables. The
measurement of observable Ω1 , is “mutually compatible” with another observable Ω2 if they
commute Ω1 , Ω2 ] = 0.

3. Eigenstates
A physical system is in an eigenstate of the operator Ω if Ω | ψn >= ωn | Ψn >, where ψn is
the nth eigenstate, corresponding to the eigenvalue n. For a Hermitian operator ωn is real.
4. Expansion Postulate - Basis
An arbitrary state vector for a physical system can be expanded in a complete orthonormal
set (CON) of eigenfunctions ψn of a complete set of commuting operators Ω1 · · · ΩN
X
Ψa = an | ψn >
n

Orthonormality < ψn | ψm >= δnm is a statement that the states are distinct. Normalization
is a statement that we have all of the state.

5. Measurements and Probability


The result of a measurement ofPa physical system is one of its eigenvalues. For a physical
system described by | Ψa >= n an | ψn > with Ω | ψn >= ωn | ψn >, measurement of a
physical observable Ω results in one of the eigenvalues ωn with the probability | an |2 . The
CHAPTER 8. SPIN 202

average of many measurements of the observable Ω on identically prepared systems is given


by the expectation value X
< Ψa | Ω | Ψa >= | an |2 ωn ,
n
i.e. that value of all the eigenvalues weighted by the probability of that eigenvalue occurring
in the state Ψa .
6. Dynamics - Time Development
The time development of a physical system is expressed by the Schrödinger equation

| Ψa (t) >= H | Ψa (t) >,
ih̄
∂t
where H is a linear Hermitian operator. The Hilbert space vector | Ψa (t) > rotates in Hilbert
space. For a closed system H has no explicit dependence on time ∂H ∂t = 0, in which case its
eigenvalues are the possible stationary states of the system. A superposition principle follows
from the linearity of H, and a statement of conservation of probability follows from the
Hermitian property of H

< Ψa | Ψa >= 0.
∂t
Properties of Complex Linear Vector Space
Reference: Schiff-Chapter 6 and Dirac Chapter 1-2.
To fully explain the meaning of the postulates we need to define the properties of the state vectors.
The vector denoted by a “ket” | Ψ > describes the state of a physical system. It is not our usual
3-D vector in space, but is an abstract notion - a vector Hilbert space. It exists in our minds. The
state vector is defined in an infinite dimensional space for the case of the coordinate (~r) and
momentum (~ p) degrees of freedom, but is a vector in a finite space for quantum mechanics degrees
of freedom such as for spin. For the ~r space case the components < ~r | Ψ(t) > of the state vector
are just the functions call wave functions - the number of components are the nondenumerable set
of points in space ~r. Often we can think of these “components” as finite - denumerable, such as we
had done earlier for box normalization. The limit as the dimension N −→ ∞ is assumed to be well
defined - a subject which requires much discussion concerning Hilbert space. To define a Hilbert
space one requires linear rules, the existence of a scalar product, and of a basis which is assumed
to be complete. In the case of square integrable functions [ d3 r | ψ |2 < ∞ or |k ψ k< ∞], the
R

completeness requires that we can get convergence “in the mean”


N
X
lim k ψ − an ψn k→ 0
N →∞
n

for “reasonable” functions ψ. I will not dwell on such matters, but simply refer you to the famous
book by J. von Neumann“Mathematical Foundations of Quantum mechanics” for formal treatment.
Addition of Vectors
Two ket vectors, which describe two possible states of the system, can be combined in a linear
fashion.
| Ψ >= a | Ψa > +b | Ψb > .
That is, we assume that superposition of states is possible which is physically required to get
interference effects. Here the coefficients a and b can, in general, be complex numbers, which is
permitted since physical quantities and probabilities are determined from bilinear products of the
above vectors.
The rules for addition are associative and commutative:
| Ψa > + | Ψb >=| Ψb > + | Ψa > commutative
(| Ψa > + | Ψb >)+ | Ψc >=| Ψa > +(| Ψb > + | Ψc >) associative) .
We shall denote these vectors by | Ψa > or just | a > . Here | a > refers to a general state of the
system, which could be some combination e of eigenstates of a complete set of commuting
operators. The state | a > is later defined to be a particular combination of such eigenstates.
e such as occurs in building wave packets
CHAPTER 8. SPIN 203

Multiplication
The rules for multiplying state vectors by complex numbers are

λ(a | Ψa >) = (λa) | Ψa >= λa | Ψa >

λ(| Ψa > + | Ψb >) = λ | Ψa > +λ | Ψb > (linear)


Here λ and a are complex numbers.
The null vector is defined by | Ψa > +0 =| Ψa >, it is required for a consistent discussion of the
general vector space, but I wish to emphasize that the null vector is not a physical state of the
system (it is not normalized to unity). Do not confuse the null vector with the vacuum state
discussed later, which is a physical state, the ground state.

Linear Independence - Basis


For N state vectors | ψ1 > · · · | ψN >, if we can not write one in terms of the others, we then have
a N-dimensional vector space of linearly
PN independent vectors. Mathematically, linear independence
requires that the only solution to n=1 λn ψn = 0 be all λn = 0 for N-space.
These N vectors, provide a basis, or a complete set, which “space the space.” Hence an arbitrary
vector in the N-dimensional vector space can be written as a linear combination of the basis vectors
N
X X
| Ψa >= an | Ψn > or | a >= an | n > .
n=1

Here an are the (complex) components of the vector | ψa > . The rules for addition and
multiplication given before now can be applied directly to the vector components (to the complex
numbers) themselves
an + bn = cn
follows from
| a > + | b >=| c >
since X X
(an + bn ) | Ψn >= cn | Ψn > .
n

Also
λan = bn
follows from X X
λ | a >= λan | n >= bn | n >=| b > .
n

Scalar Produce
~·B
We now define a scalar or dot product, in analogy to the usual scalar produce A ~ = Ai Bi . First
we associate with each ket, a dual (or adjoint) vector

a | Ψ >−→ a∗ < Ψ | or < Ψ | a∗

Also written as < aΨ |=< Ψ | a∗ .


The scalar product is then defined by

< Ψa | Ψb >≡< a | b >=< b | a >∗ =< Ψb | Ψa >∗

< λΨa | Ψb >= λ∗ < Ψa | Ψb >


or
< λa | b >= λ∗ < a | b > .
We also have
< Ψa | Ψb + Ψc >=< Ψa | Ψb > + < Ψa | Ψc >
or the equivalent
< a | b + c >=< a | b > + < a | c > .
CHAPTER 8. SPIN 204

Finally note that:


< Ψa | Ψa > > 0
or in the alternate notation
< a | a > > 0.
holds true for all states, except for the null, unphysical state.
The norm of the state is defined by < a | a >, which is taken to be 1

< Ψa | Ψa >= 1.

Furthermore, the basis vectors are also normalized and orthogonal. If orthogonal ⇒ linearly
independent.
< n | n0 >= δnn0
The above properties assure that we can interpret the real positive probability p(a b) and the
complex probability amplitude < a | b > in a meaningful way:

p(a b) ≡|< a | b >|2 = real and > 0.

See later for more discussion.


Note in general < a | b >6= δab since they are general states not basis states.
The expansion of the general states | a > and < b | now leads us to

X N
X
< a | b >= a∗n < n | n0 > bn0 = a∗n bn
n,n0 n

N
X N
X
< a |= a∗n < a |←| a >= an | n > .
n n=1

Here the dot product takes the usual form in terms of components, except we need a∗n to apply to
complex space. Occurrence of a∗n , which arises from the dual space property given earlier, leads to
N
X
< a | a >= | an |2 = 1 > 0
n=1

PN
Application of these rules to | a >= n=1 an | n > gives us an expression for the components in
dot product form
< n | a >= an ; a∗n =< a | n >
Using this result, our previous expressions become simply:
X
| a >= | n >< n | a >
n
X
< a |= < a | n >< n |
n
X X
< a | b >= < a | n >< a | b >= a∗n bn
n
X X
< a | a >= < a | n >< a | a >= | an |2
n
X X

<a|b> = b∗n an = < b | n >< n | a >=< b | a >
n n
PN
All of which is summarized by n=1 | n >< n |= 1 = unit operator!
CHAPTER 8. SPIN 205

8.6.1 Operators
Now we go on to aspects of the second postulates; namely to define operators in the complex
vector space. An operator changes one vector to another

| Ψ 0 >= A | Ψ >, (8.11)


f
where A is an (abstract) operator in the vector space. To make it a linear operator , we need to
satisfy both
A(| Ψa > + | Ψb >) = A | Ψa > +A | Ψb >, (8.13)
and
Aa | Ψa >= a A | Ψa > . (8.14)
Some additional rules for operators are: if A | Ψ >= B | Ψ >, for all | Ψ >, then A ≡ B. ( Note we
must be careful to recognize operator equations as opposed to matrix element equations.)

(A + B) | Ψ > = A | Ψ > +B | Ψ > (8.15)


A (B | Ψ >) = (A B) | Ψ >,

but
A B | Ψ >6= B A | Ψ > (8.16)
... in general operators do not commute
[A, B] 6= 0, (8.17)
which is called a non-Abelian group property.
Finally, the unit operator is defined by

I | Ψ >=| Ψ >, (8.18)

for all | Ψ >’s.


The closure expression is now understood as a unit operator
X
| n >< n |= I = unit operator (8.19)
n

f An antilinear operator, used to represent time reversal or reversal of motion, is defined by the first property above
and
à a | Ψa >= a∗ à | Ψa > . (8.12)
à can be represented by à = A K0 , where A is a linear operator-acts only on the vector and K0 acts only to change
i to −i, wherever i appears.
CHAPTER 8. SPIN 206

8.6.2 Projection operators


The closure equation shows that a unit operator I can be constructed from the bra’s and ket’s.
Another example of using Hilbert space vectors to construct an operator is the projection operator

Pn =| n >< n | .

Clearly, X
Pn = I,
n

and
Pn Pm =| n >< n | m > < m |= δnm Pn ,
which defines a projection or idempotent operator.
We can also apply Pn to general states

Pn | Ψa >=| n >< n | Ψa >=| n >< n | a > .

Here Pn projects out the nth component < n | a >= an .

Matrix Representation of Operators


Now we can make the idea of operators less abstract by applying the projection operators, using a
basis | n >, to obtain a matrix representation of operators. The steps are for an operator A
X
A | Ψa >≡ A | a >= A | n0 >< n0 | a >=| Ψb >≡| b > .
n0

Using closure again we get


X X X
A | a >= | n > < n | A | n0 >< n0 | a >= |n><n| b>.
n n0 n

Since the basis | n > is CON, we can identify the operator as


X X X X
A= | n > < n | A | n0 >< n0 |= Pn APn0 .
n n0 n n0

We also see that the final vector | b > produced by the action of A on the initial state | a > can
now be expressed in terms of the components
X
< n | b >= < n | A | n0 >< n0 | a >,
n0

or
N
X
bn = Ann0 an0 :
n0

which is a matrix equation b = Aa. The above equations involve only complex number in the form
of column N × 1 matrices for a  
a1
 a2 
an →  . 
 
 .. 
aN
and for b  
b1
 b2 
bn → 
 
.. 
 . 
bN
CHAPTER 8. SPIN 207

and a N × N matrix for A


 
A11 A11 ··· A1N
 A21 A22 ··· A2N 
Ann0 →  .
 
.. .. .. ..
 . . . . 
AN 1 AN 1 ··· AN N

We have replaced the operator in the Hilbert space by a N × N matrix equation, which we call the
matrix representation. However, the N −dimensional space need not be finite or discrete (as it is
for spin), but could be infinite and continuous (as it is for r− or p− space). Recall that the
operator p~ could be represented in matrix form as
h̄ ~
(~r | p~ | ~r 0 ) = ∇~r δ(~r − ~r 0 ),
i
as a continuous, infinite, off diagonal matrix. g
The full power of the Hilbert space formulation and of Dirac’s notation is seen here. we can
express finite matrices, differential operators all on the same footing. Also functions and matrices
are simply related, which serves to unify wave mechanics and matrix mechanics into a single
theory, called Dirac’s transformation theory or simply Quantum Mechanics.

More on Matrices
For two operators, the sum A + B = C, clearly maps over to the addition of the matrices, provided
they are all of the same dimensionality (N × N ) :

< n | A | n0 > + < n | B | n0 >=< n | C | n0 > .

For multiplication AB = C, we see that the usual matrix multiplication follows from
X
C | a >= AB | a >= |b >= bn |n >
n
X X
= | n >< n | AB | n0 >< n0 | a >= | n >< n | C | n0 >< n0 | a >,
n n

hence we get the usual rule for matrix multiplication


X
< n | C | n0 >=< n | AB | n0 >= < n | A | n00 >< n00 | B | n0 > .
n00

8.6.3 Hermitian Conjugate or Adjoint Operator


We have introduced a dual space by associating < Ψb | with each ket | Ψb > . Now we introduce
operators acting first on just these adjoint state vectors

< Ψa | A† =< Ψb |

i.e. A† takes < Ψa | and makes it < Ψb | in the adjoint space. Our closure property (or the use of a
complete orthonormal basis) permits us to make the above equation into a matrix one
X X
< a | A† | n0 >< n0 | = < b | n0 >< n0 | ,
n0 n0

or X X
b∗n0 = < a | n >< n | A† | n0 >= a∗n < n | A† | n0 > .
n n

Using complex conjugation and the properties a∗n =< a | n >=< n | a >∗ , we obtain from the
above equation X X
bn0 = < n | A† | n0 >∗ an = < n0 | A | n > an ,
n n
g There are pitfalls in this representation because of the special nature of singular functions.
CHAPTER 8. SPIN 208

therefore
< n0 | A | n >=< n | A† | n0 >∗
which is the usual definition of the adjoint operator; namely the transpose T and conjugation ∗

A† = AT ∗

It is not just for the (CON) basis states | n > that the above property holds, but for any general
states < b | and | a > . The proof is:
X
< a | A† | b > = < a | n >< n | A† | n0 >< n0 | b > (8.20)
n n0
X
= ( < b | n >< n0 | A | n >< n0 | a >)∗
n n0
= < b | A | a >∗ ,

which is again a transpose and a conjugation, but now on general states.


Finally, we associate with | b >= A | a >, the adjoint state, < b |=< a | A† =< A a | . Hence

< A a | b >=< a | A† | b >,

also defines the adjoint operator. We saw this before as


Z Z
Ψ∗a A† Ψb dτ = (AΨa )∗ Ψb dτ,

which is an integration by parts step for the r− and p− space representations.

Hermitian Operators
Self-adjoint operators are called Hermitian and are used to represent observables (postulate #2) In
this section, we use the symbol Ω to denote a general Hermitian operator

A† = A = Ω.

These represent observables because their expectation value is real


X
< a | Ω | a >=< a | Ω† | a >=< a | Ω | a >∗ = | an |2 ωn ,
n

where we took Ω | n >= ωn | n > . It follows that the eigenvalues are real for a Hermitian operator.
We can now interpret | an |2 as

| an |2 =< a | n >< n | a >= p(n, a),

where p(n, a) is the probability of finding (or measuring) the eigenstate ( and its associated
eigenvalue ωn ) in the general state | a > . This qualifies as a probability since it is real and sums to
one if all possibilities are included
N
X X X
p(n, a) = | an |2 = < a | n >< n | a >=< a | a >= 1.
n=1 n n

The positive quantity gives the probability of finding the system, which was prepared in the
general state | a >, with the eigenvalue ωn . The sum over all possible situations of course leads to a
total probability of unity (see Postulates 3-5).
CHAPTER 8. SPIN 209

Inverse Operators
The inverse operator is defined as follows: if | b >= A | a > then the inverse operator reverses this
transformation
| a >= A−1 | b >= A−1 A | a >,
for all states. Then
A−1 A = I
or X
< n | A−1 | n00 >< n00 | A | n0 >= δnn0 .
n00

A solution exists only if det A 6= 0. If that is satisfied then we have an inverse operator and can
define a similarity transformation

Similarity Transformations

A similarity transformation generates new vectors | g


n >, by taking linear combinations of the
original basis vectors | n >,
X
|g
n > = S | n >= < n0 | S | n > | n0 > .
n0

Recall that the original basis vectors | n > are a complete orthonormal set < n | n0 >= δnn0 . Since
the inverse S −1 is assumed to exist, we can also use | g
n >, to construct the basis | n >,

| n >= S −1 | g
n >.

The new vectors | g


n > form a complete set of states, i.e. they are a linearly independent set, which
can be used to construct general states. We have the general state given by
X
| a >= < n | a > | n >,
n

and hence our new states | g


a > are
X
S | a >= | g
a>= < n | a > |g
n >,
n

where | g
n > is used to get a general state | g
a >.
The significance of a similarity transformation is that it yields the same eigenvalues for the
appropriate transformed operator. We have

Ω | n >= ωn | n >

and
(SΩS −1 )S | n >= ωn S | n >
or
Ω̃ | g
n > = ωn | g
n >.
Thus the transformed operator Ω̃ ≡ SΩS −1 has the same eigenvalues as the original operator Ω,
but with new eigenfunctions | gn >.
If we also require that the new set be not only complete, but also orthonormal, we must add the
restriction that
<g 0 > =< n | S † S | n0 >= δ 0 ,
n |ng nn

for all n, n0 . We arrive then at the condition

S†S = I

that S is a unitary operator, or equivalently that S −1 = S. Similarity transformations that also


preserve orthonormality are therefore unitary operators and will will denote them by U.
CHAPTER 8. SPIN 210

Unitary Operators - Canonical Transformations


Unitary operators play an important role in quantum mechanics. They provide a means of
changing our description using alternate, often more convenient, basis states. In classical physics
corresponding changes in our choice of generalized coordinates, defined to give a change in
description, are called canonical transformations. In classical physics, such transformations reveal
the basis symmetries of the dynamical system - as will be the case for the quantum theory of
canonical transformations.
Unitary operators serve to represent canonical transformation or changes of our descriptions in
quantum theory. Later we will discuss the transformation for translations, rotations and
time-evolution. This will lead to the study of unitary groups and symmetries.
The change in description is done by transforming from one complete orthonormal set to another
complete orthonormal set:

|g
n>=U |n> U †U = 1

0 > =< n | n0 >= δ 0 . Also from closure on the original basis


n |ng
with <g nn
X
| n >< n |= I,
n

we deduce X
U | n >< n | U † = U IU † = I,
n

or closure on the new states X


ng n | = I, .
><g
n

n > form a new CON set. h The action of U on general states | a >, | b >
Thus we see that | g
leaves their normalization unchanged since |g
b > = U | b > and

b > =< b | U † U | b >=< b | b >= 1.


gb|g
<

In addition the overlap < a | b > remains unaffected by the unitary transformation

b > =< a | U † U | b >=< a | b >= 1.


ga|g
<

Here we apply the same U to all states | a >, | b > . Hence the probability
p(ab) =|< a | b >|2 =| <
ga|gb > |2 is unchanged by a unitary transformation. That is why we call it
a canonical transformation,; U produces a change in description without changing the predictions
as expressed by the probabilities.
Earlier we saw that a similarity transformation leaves eigenvalues unchanged, now we see that a
unitary operator has the same property plus probability amplitudes are unaltered. For example,
we have

Ω | n >= ωn | n > ↔ (U ΩU † ) | g
n > = ωn | g
n >.

The transformed operator has the same spectrum or eigenvalues ωn when acting on the new CON
set | g
n >.
For a general state, the expectation value
X X
< a | Ω | a >= |< a | n >|2 ωn = p(n, a) ωn .
n n

where p(n, a) is the probability of finding an eigenvalue ωn in the state | a >, with ωn being the
result of a particular measurement.
h From U † U = I, it is easy to deduce that U U † = I.
CHAPTER 8. SPIN 211

After application of a unitary operator we get the same value for the expectation values

<g a > = < a | (U † U )Ω(U † U ) | a >=< a | Ω | a >


a |Ω̃| g (8.21)
X X
= |< ga| g n > |2 ω̃n = p̃(n, a)ωn .
n n

We see that the eigenvalues ω̃n = ωn are equal and the probabilities are equal p̃(n, a) = p(n, a)
ga| g
since < n > =< a | n > .
The significance of these features of unitary operators is that no change in physics is generated -
the probabilities, eigenvalues, and, hence, all expectation values are unchanged. Hence we can
consider the application of a unitary operator on all states and operators to be simply a change in
description or a canonical transformation.
An alternate way to state this property is that U leaves both the lengths <ga| g
a > =< a | a > . and
ga| b > =< a | b > . for all vectors in the space unchanged; hence, we have simply
angles < g
performed a rotation in complex vector space - or chosen a new basis for our description.
Examples of the form for unitary operators are

U = eiΩ

which is unitary U † U = I, if Ω is Hermitian Ω† = Ω. Another possible form is


I + iΩ
Ω=
I − iΩ

which is also unitary U † U = I, if Ω is Hermitian Ω† = Ω.


Also note that a product of unitary operators is still a unitary operator. U = U1 U2 satisfies

U † U = U2† U1† U1 U2 = 1,

which is a property of a group.


Unitary operators can also be written in the following form in terms of the two CON sets.
X X
U= |g
n > < n |= U | n >< n |= U I.
n n

or more generally
CHAPTER 8. SPIN 212

8.6.4 The Algebra of Measurement - A Brief Visit


References: Schwingers lectures Quantum Kinematics and Dynamics ,Gottfried Chapter V, and
Sakurai Modern Quantum Mechanics.
Schwinger formulates quantum theory by expressing the properties of microscopic measurement in
a symbolic manner using measurement operators. He does not begin by listing postulates and
assuming a linear vector space. Instead he expresses the basic property of microscopic
measurements, especially the limitations contained in the uncertainty principle, at the first stage
and then induces a measurement algebra and a vector space representation from an analysis of the
physical properties of measurement. It is an elegant approach and I wish we had the time to
discuss it fully. We do not. Hence I will merely outline the basis points. I strongly urge you to look
at Schwingers notes (and Gottfrieds book and Sakurais book) for their discussion.
Since the interaction between a measuring apparatus and an atomic system can not be made
arbitrarily small (as one can to a good approximation for classical systems), nor can it be
compensated precisely - the interaction between system and instrument is to some extent
uncontrollable and unpredictable. (This is a statement of the uncertainty principle, which applies
to small systems i.e. of small action). A measurement of one property can produce unavoidable
changes in the value of a previously assigned property. An assignment of precise values for all of
the attributes of a microscopic system can not always be made.
Schwingers statement “ ” says it so well.
The simplest measurement process (as for example given by the Stern-Gerlach apparatus) involves
sorting an ensemble of similar systems into sub-ensembles distinguished by definite values
(1) (2) (3) (n)
a1 , a1 , a1 , · · · a1 of a physical quantity A1 . ( Here we adopt Schwinger’s notation for what we
earlier designated as an observable Ω and its eigenvalues ωi .) Now use a prime to denote one of the
(1) (2) (3) (n)
values associated with observable A1 , i.e. a01 is one of the set a1 , a1 , a1 , · · · a1 , The symbol
0
M (a1 ) is then introduced by Schwinger to symbolize the selective measurement associated with A1
wherein only systems with particular value a01 of attribute A1 are accepted. For example, in the
SGE A1 refers to the magnetic moment of an atom or its spin and a01 would then be one values of
ms . either +1/2 or −1/2. The picture is then

@ S
@
M (+1/2)
p p p pI
p p p pp -
p p p p p p p p p p p pp pp p p p p p p p p p p p p p p s a01 = +1/2
@
p p p p
p p p p
@
R @
p p p p
p ps 0
N a1 = −1/2
Oven Atomic Beam Magnet

Figure 8.10: The Stern-Gerlach Experiment (SGE) setup, where the measurement symbol M (+1/2)
represents the selection of the spin up systems or sub-ensembles.

where the spin up direction is drawn along a straight line as done by Schwinger to help in the
induction process. The sorting of the incident ensemble into the two classes (for the case of spin
1/2) is the symbolized by the two “operators” M (ms = 1/2) and M (ms = −1/2). Then
mathematical steps associated with the measurement are induced i as follows:
1. Addition
the sum of these symbolsM (a01 ) + M (a001 ) represents that both a01 and a001 systems are
accepted.
i Schwinger points out this is and inductive rather than a deductive procedure
CHAPTER 8. SPIN 213

2. Multiplication
M (a001 )M (a01 ) represents a sequence of two SG measurements in order; first select a01 then a001 .
That is we order the symbols with the convention: first measurement symbol to the right and
the next to the left.
Physically we see that two sequential SGE’s which select the same attribute (eigenvalue) selects a
value already selected so we have the symbolic relation

M (a01 )M (a01 ) = M (a01 ).

Alternately, if we first do a SGE to select a01 followed by a SGE which selects a different possible
value a001 of the attribute A1 , then
M (a001 )M (a01 ) = 0,
if a001 6= a01 . The first SGE already rejected the a001 sub-ensemble.
On the other hand if a single SGE is performed with all cases accepted we have the relation
X
M (a01 ) = 1,
a01

which means for the spin up/down case we accept both the spin up and down possibilities. For the
case of silver deposits it would mean we collect all of the silver in both branches. This is
completeness. We also see that the ”symbol’ 1 denotes all cases accepted, whereas the ”symbol’ 0
denotes all cases rejected. Schwinger then gives physical meaning to the “symbolic” steps of

1×1=1 1×0=0×1=0

and
0M = 0 1M = M,
etc. All of this might seem mysterious. It is simply expressing the ability to select (for SG,
discrete) sub-ensembles with definite attributes, such as occurs in the SGE. Indeed, the
measurement symbol M (a01 ) is simply given by a projection operator j

M (a01 ) ≡| a01 >< a01 |,

but that is not induced until much later in Schwinger’s approach. Nevertheless, much of his
beautiful formulation can be understood if we keep the above “secret” in mind. Clearly, since

| a01 >< a01 | a001 >< a001 |= δa01 ,a001 | a01 >< a01 |,

and X
| a01 >< a01 |= 1,
a01

the above properties for M (a1 ) follow from our earlier discussion of CON bases.
If other simultaneously observable attributes of a second observable A2 are measured, then the
measurement symbol is extended to include these additional attributes of the sub-ensembles. We
have the extension of the measurement symbols to:

M (a02 , a01 ) = M (a02 )M (a01 ) = M (a01 )M (a02 ) = M (a01 , a02 ), (8.22)

which states the measurement result of a01 of attribute A1 followed by a measurement of a02 of
attribute A2 can be done in reverse order with no change in the results. That is, the attributes are
compatible or simultaneously observable and the order of the SGE commute. The above property
for the case of measuring compatible observables can also be expressed as

(| a02 >< a02 |)(| a01 >< a2 10 |) =| a01 a02 >< a01 a02 |≡ M (a1 , a2 ) = M (a), (8.23)
j We had earlier denoted the projection operator as Pn =| n >< n | .
CHAPTER 8. SPIN 214

where in the final M (a) the designation a denotes a full set of such compatible observables
a → a1 , a2 , · · ·. For example, we could have a1 → s2 , a2 → ms as compatible spin observables or
a1 → energy, a2 → `2 , for atomic levels.
The idea of compatible observables is: “Two physical quantities (A1 , A2 ) are said to be compatible
when the measurement of one does not destroy the knowledge gained by previous measurement of
the other.” the order of the measurements can therefore be switched as done above. Here we deal
with the commutation of the observables [A1 , A2 ] = 0 or for the measurement operators M ,
[M (a1 ), M (a2 )] = 0. The idea of a product space is based on this physical idea of compatible
observables and the product of the measurement operators, where their representation in
projection operator form allows one to connect the product space idea to the product of
commuting measurement symbols.
Going beyond mutually compatible ( simultaneously measurable) observables, Schwinger states ”A
more general type of measurement incorporates a disturbance that produces a change of state.”
Such cases are are represented by non-commuting measurement operators.
The complete set of compatible physical attributes A : A1 , A2 · · · AN , for which [Ai , Aj ] = 0 for all
values of i and j, are not the only possible attributes of the system. Another set of compatible
attributes B : B1 , B2 · · · BN , for which [Bi , Bj ] = 0 for all values of i and j, can also be considered.
For example, the set A could refer to the quantum operators H, s2 sz , `2 , `z , whereas another set B
could be the H, `2 , `z , j 2 jz compatible operators. In view of this possibility, one can consider
measurements for which only systems with value b0 of B are let in, but only systems with values a0
of A are left out. This is represented by a measurement symbol M (a0 | b0 ). Note here the primed
quantities are the measured set of values as in b0 − > b01 b02 · · · b0N . Jumping ahead to the
representation of the properties of such measurement as analyzed by Schwinger, one obtains the
connection
M (a0 | b0 ) =| a0 >< b0 | . (8.24)
Many properties can be seen to follow from the measurement process, which is the procedure used
by Schwinger to develop the algebra of the measurement symbols M (a0 , b0 ). The idea of an adjoint,
the physical meaning of a trace are all described in his book. Only after a full development of such
properties does the above representation in terms of the Hilbert space operator | a0 >< b0 | emerge.
I will not pursue this direction further. If you keep in mind that M (a0 | b0 ) =| a0 >< b0 |, then you
should be able to follow and appreciate the discussions given by Schwinger and also by Gottfried
and Sakurai.
Chapter 9

Symmetries

References:
Unitary operators permit us to change from one complete orthonormal (CON) basis to another
CON basis. Therefore, we can change our description without altering the eigenvalues or the
probabilities. Hence the expectation values are unchanged when a unitary operator is applied to all
states and all operators. in the following manner

a |( U † A U )|g
< a | A | b >= <g a |Ã|g
b > = <g b >.

Now let us consider explicit cases of unitary transformations, which represent a change of
description. These are called canonical transformations and the cases we will consider will be
changes due to translation or rotation of our laboratory and later, time development.
The unitary operators will be a means of making kinematical transformations ( or a change of
description). The invariance of the equations of motion under certain kinematical transformations
will lead us to conservation principles. In this way, symmetries of the system are seen to play an
important role. We shall see that conservation of linear momentum is related to the symmetry or
invariance of the equations of motion under spatial displacement. Similarly, rotational symmetry
leads to conservation of angular momentum and displacement in time invariance leads to
conservation of energy. These general remarks, which are valid in both classical and quantum
mechanics (see Goldstein) will be shown in our subsequent discussion.
Let us begin with the simplest case-that of spatial translation or displacement.

9.1 Spatial Translation


Consider a laboratory ( containing a beam, an accelerator, lasers, detectors and people, i.e.
everything involved in making a measurement) which is located in the region “LAB” shown below.

LAB LAB
g

-
∆~x
Ψ(~r1 , ~r2 , ~r3 , · · · , ~rN , t) e r1 , ~r2 , ~r3 , · · · , ~rN , t)
Ψ(~

Figure 9.1: The active translation of a LAB to a new site LAB ~ applied to all
g by a displacement ∆x
N particles.

The laboratory, and the experiments performed in it, are describes by a state vector, or a wave
function”
Ψ(~r1 , ~r2 , ~r3 , · · · , ~rN , t),
CHAPTER 9. SYMMETRIES 216

which describes the N −particles making up the whole laboratory. Clearly,Ψ(~r1 , ~r2 , ~r3 , · · · , ~rN , t), is
large within the laboratory region and small away from it, since that is where the laboratory
is–isn’t it?
Now let us carry out anactive change in description; namely, let us move (at NSF expense) the
entire laboratory to a new location LAB
g by translating it by ∆x.~ Every particle has been
displaced by the same amount. Here we take the active transformation and move the laboratory
LAB → LAB, g we shall stick with this active approach. The alternative is to describe the
translation by the passive transformation of simply moving the coordinate frame, which is clearly
equivalent except one translates the frame by −∆x.~
Once the entire laboratory is located at LAB,
g the wave function describing all of the laboratory
(beam, accelerator, lasers, detectors and people, i.e. everything involved in making a
measurement) is now given by a new function

e r1 , ~r2 , ~r3 , · · · , ~rN , t),


Ψ(~

of the old coordinates ~r1 , ~r2 , ~r3 , · · · , ~rN . The new function is simply one that is nonzero in the
g and is nonzero outside of that region. a
region of LAB
Now we introduce an operator to represent the act of translating the system by considering the
operator’s action on the original wave function to yield the above new function

U Ψ(~r1 , ~r2 , ~r3 , · · · , ~rN , t) = Ψ(~


e r1 , ~r2 , ~r3 , · · · , ~rN , t),

where Ψe is a function of the old coordinates. From the previous discussion of the act of translating
the laboratory in such a way as to restore it to the same relative setup, it follows that

~ ~r2 −∆x,
e r1 , ~r2 , ~r3 , · · · , ~rN , t) = Ψ(~r1 −∆x, ~ ~r3 −∆x,
~ · · · , ~rN −∆x, ~ P
~ t) = e− h̄i ∆x· ~
Ψ(~ Ψ(~r1 , ~r2 , ~r3 , · · · , ~rN , t).

To reconstitute the laboratory at LAB


g just the way it was at LAB, the new function of the old
coordinates is set equal to the old function of shifted coordinates. The last step introduces an
exponential, which arises from expressing a Taylor series expansion as an exponential of a
derivative, i.e.

X 1 ∂
f (x − ∆x) = e−∆x ∂x f (x) = (−∆x)n ( )n f (x).
n
n! ∂x
When applied to all coordinates, this yields
~
Pj=N ~j
~ ~r2 − ∆x,
Ψ(~r1 − ∆x, ~ ~r3 − ∆x, ~ t) = e−∆x·
~ · · · , ~rN − ∆x, j=1

Ψ(~r1 , ~r2 , ~r3 , · · · , ~rN , t).

Therefore, the linear momentum operator is identified


j=N
i X~
P~ = ∇j .
h̄ j=1

We conclude that the operator which changes the description of our system (from being at LAB to
being at LAB
g is
i ~ ~
U = e− h̄ ∆x·P ,
with P~ being the total momentum operator given above. Since P~ † = P~ , it follows that

U † U = U U † = 1,

so we indeed have a unitary operator. The CON basis defined in LAB has been replaced by the
translated CON basis in LAB.
g
a We take all of this at one common, particular instant of time.
b Note that we introduced h̄ here, but the result for the translation operator U is independent of h̄ and holds for
classical translations.
CHAPTER 9. SYMMETRIES 217

~ P
~ ≡ e− h̄i ∆x· ~
Additional properties of U (∆x) are
~ = U −1 (∆x)
U † (∆x) ~ = U (−∆x),
~
~ translation. Also, the no translation case
an inverse operator exists and is related to a −∆x
U (0) = 1,
and hence a unit operator exists. Finally, two sequential translations yield a translation
~ 1 )U (∆x
U (∆x ~ 2 ) = U (∆x
~ 1 + ∆x
~ 2 ),

thus these unitary operators form a group, the translation group. Indeed, these properties show
that we are dealing with a Lie group, a continuous unitary group. Finally we note that
e b >=< Ψa | U † U | Ψb >=< Ψa | Ψb >,
ea | Ψ

for two general states (labelled a and b); hence, U preserves the normalization and orthogonality of
the basis.
Now consider the energy eigenvalue E, assuming that | Ψ > had a definite energy. c
H | Ψ >= E | Ψ >,

then with H̃ = U HU , we have
H̃ | g
Ψ > = E| g
Ψ >,
which means that the new wave function Ψ gives the same eigenvalue provided the new
e
Hamiltonian H̃ is used. In general H̃ will be a new function of the old coordinates ~r1 · · · ~rN . and
momenta p~1 · · · p~N . If for example the LAB is moved into a region with strong electromagnetic
fields that were not in the original location, then to get the same eigenvalue E, terms would have
to be included in H̃ to compensate for these fields and H̃ would be a new function. Application of
H to | gΨ > yields a new energy E 0 = <g Ψ |H | g
Ψ > 6= E, in general.
If the new and old Hamiltonians are equal, i.e. the same function of ~r1 · · · ~rN . and p~1 · · · p~N , then
we say there is an invariance of the equations of motion under spatial translation and it follows
that H̃ = H; hence
Ψ |H̃ | g
E = <g Ψ |H | g
Ψ > = <g Ψ > =< Ψ | H | Ψ > .
Since H̃ = H, it follows from the invariance of the equations of motion, or from translational
symmetry that total linear momentum P~ is conserved
dP~
[P~ , H] = 0 or = 0.
dt
To see this more clearly, consider the case of an infinitesimal translation ∆~x → 0 where the new
Hamiltonian H̃ is generated from the old on H by
i ~ i ~ i
H̃ = U H U † = e− h̄ ∆~x·P H e+ h̄ ∆~x·P ' H − ∆~x · [P~ , H] + O((∆~x)2 ) · · ·

For H̃ = H, this yields [P~ , H] = 0.
Another way to express this invariance is to use the full (non-infinitesimal) unitary transformation
U H − HU = [U, H] = 0,
which follows from H̃ = H = U H U , and then HU = U H U † U = U H.

This invariance holds only if the situation at LAB and LAB g are the same, in the sense that no
special properties, such as a strong electric field exits in only the new LAB g region. In general the
invariance of H under spatial translation is simply a statement of the uniform nature of space.
This description is part of Emmy Noether’s theorem. See the courseweb links for biographical
information about her. d It is worthwhile to read about her, especially the tribute to her by
Einstein.
c We could take HΨ = ih̄ ∂ Ψ with no substantial change in our discussion here. We would also have H̃ Ψ ∂ e
e = ih̄ ∂t Ψ,
∂t
with H̃ generating the changes in Ψ
e , while H generates the changes in Ψ for t > 0.
d Here are some of the classic references: Invariante Variationsprobleme,” Nachr. v. d. Ges. d. Wiss. zu Gttingen

1918, pp 235-257; E. Noether, Nacht. kgl. Ges. Wiss. Gottinger, 171 (1891); Hill, E. L. Rev. Mod. Phys. vol 23,
1951, p253; Havas, P. and Stachel, J., Phys. Rev. 185, 1969, page 1636; Boyer, T., Am. J. Phys, v 34, 1966 p475.
CHAPTER 9. SYMMETRIES 218

9.2 Rotations
Consider our LAB, which as before includes everything, to be rotated at a given time t by an
angle α in the positive direction about the ẑ axis as shown in the figure. This is an active rotation
in that we move the laboratory.


C
 C
CC LAB
g C
 e r1 · · · , ~rN , t)
: C Ψ(~ C
  C C

 C    C
   C  
 α
 -
Ψ(~r1 · · · , ~rN , t)
LAB

Figure 9.2: The active rotation of a LAB to a new site LAB


g by an angle α about the ẑ axis applied
to all N particles.

Again, Ψ(~r1 , ~r2 , ~r3 , · · · , ~rN , t) describes all N particles in the original laboratory; therefore Ψ peaks
mostly in the region LAB and is zero outside.
Now we represent the act of rotating our system, i.e. of changing the description, by an operator D
acting on Ψ. we have

D Ψ(r1 θ1 φ1 , r2 θ2 φ2 , r3 θ3 φ3 , · · · , rN θN φN , t) = Ψ(r
e 1 θ1 φ1 , r2 θ2 φ2 , r3 θ3 φ3 , · · · , rN θN φN , t),
e 1 θ1 φ1 , r2 θ2 φ2 , r3 θ3 φ3 , · · · , rN θN φN , t) is a new function of the old coordinates which is
where Ψ(r
peaked in the region where the laboratory has been moved to LAB. g Using steps similar to the
translation case, we have

e 1 θ1 φ1 , r1 θ1 φ1 , r1 θ1 φ1 , · · · , rN θN φN , t) = Ψ(r1 θ1 φ1 −α, r2 θ2 φ2 −α, r3 θ3 φ3 −α, · · · , rN θN φN −α, t),


Ψ(r

the new function of the old coordinates is set equal to the old function Ψ of shifted coordinates.
That is a statement that the move has restored the laboratory in all of its detail except it is
relocated by a common angular change of α about the ẑ axis.
Using the Taylor series represented as an exponential of a derivative

D Ψ(r1 θ1 φ1 · · · , rN θN φN , t) e 1 θ1 φ1 · · · , rN θN φN , t)
= Ψ(r (9.1)
= Ψ(r1 θ1 φ1 − α · · · , rN θN φN − α, t)

−α( ∂φ +··· ∂φ∂ )
= e 1 N Ψ(r1 θ1 φ1 · · · , rN θN φN , t)
− h̄i α Lz
= e Ψ(r1 θ1 φ1 · · · , rN θN φN , t).

where
N
X h̄ ∂
Lz ≡ .
n=1
i ∂φn
Here we follow the same procedures as for translations. The new function clearly peaks in the
region LAB
g and is zero outside. In LAB g the arrangement of equipment is the same as it was in
LAB at the same time. Hence Ψ can be written in terms of Ψ with shifted coordinates. the
e
Taylor’s series is again used and we find the unitary operator
i
D(α) = e− h̄ α Lz
,
CHAPTER 9. SYMMETRIES 219

with L†z = Lz =
P
n `nz . The following properties hold

D† (α) D(α) = I = D(α) D† (α),

it is unitary; a unit operator exists D(0) = I; and so does an inverse D−1 (α) = D† (α) = D(−α).
Also, for just the ẑ rotations we have a group property, D(α2 ) D(α1 ) = D(α1 + α2 ) which will be
generalized in several ways.
Generalization to rotation by an angle α about a general axis stipulated by a unit vector n̂, with
(nx , ny , nz ) = (sin θ cos φ, sin θ sin φ, cos θ), is
i ~
D(α, θ, φ) = e− h̄ α (n̂·L) ,
which requires that three numbers be specified to describe the rotation. These angles are α and
the two directions for n̂. Later we shall use three Euler angles to account for the property that as
one builds up finite rotations from infinitesimal rotations, the axis of rotation changes direction (
see later).

9.2.1 Spin and Total Angular Momentum


So far, we have generated only the orbital angular momentum operator; let us now include spin.
The reason only L~ has been obtained is because we dealt just with the spatial part of the wave
function. Now we tack on the spin part for all N particles in the LAB frame

Ψ = Ψ(r1 θ1 φ1 · · · , rN θN φN , t) | s ms >1 · · · | s ms >N .

Under a rotation about the ẑ axis, the spin states change only in their phase. The phase change is
given as:
σz
ms >1 = D | s ms >1 = e−iα 2 | s ms >1 = e−iαms1 | s ms >1 ,
| sd
for each particle with spin 1/2. This phase change is similar to the eimφ change that occurs in
Y`m (θ, φ) with eim(φ−α) → eimφ e−imα . This choice of phase for spin is made to yield proper
addition of spin to the orbital angular momentum.
Applied to all N − spins, we get D | s ms >1 · · · | s ms >N = e−iα Sz | s ms >1 · · · | sms >N ,
with
XN
Sz = szn ,
n=1

which is the total spin of the system projected on the ẑ direction. e The generalization to include
spin and a general direction by an angle α around an axis n̂ is therefore
~
Ψ > = e−iα(n̂·J) | Ψ >,
D | Ψ >= | g

with
J~ = L
~ + S,
~
N
X
~ =
L ~`n
n=1

and
N
X
~=
S ~sn .
n=1

We shall consider the order in which one performs rotations later. We will also see how this
discussion leads to angular momentum commutators.
e We assume for the purpose of this discussion that the system is made of fermions, such as electrons and nucleons.
CHAPTER 9. SYMMETRIES 220

9.2.2 The Rotation Group


~
The operator D = e−iα(n̂·J) is unitary and serves to describe an (active) rotation of our system,
including spin. The three parameters (α plus the two angles in n̂) stipulate the particular rotation;
later we shall use the three Euler angles called α, β, γ to specify such a rotation. For now let us call
these angles {α, β, γ} ≡ R, where R denotes a specific rotation. We have the unitary property

D† (R) D(R) = D(R) D† (R) = I,

the existence of an inverse


D† (R) = D−1 (R) = D(R−1 ),
where R−1 ≡ {−γ, −β, −α}, simply reverses the rotation with the proper order to return the
system to the original position. We also have the unit operation

D(0) = I,

and the group property


D(R2 ) D(R1 ) = D(R3 ),
where we will say more about this process later.

9.2.3 Invariance under Rotation


i ~
The unitary operator D = e− h̄ α(n̂·J) describes a rotation which is a change in description or a
kinematical transformation. We now apply D to the Hamiltonian and to the equation of motion in
each frame
∂ i 0
H | Ψ >= ih̄ |Ψ> or Ψ(t0 ) = e− h̄ H(t −t) Ψ(t)
∂t
in LAB and
∂ g e 0 ) = e− h̄i H̃(t0 −t) Ψ(t)
H̃ | g
Ψ > = ih̄ |Ψ> or Ψ(t e
∂t
in LAB.
g The transformed state and Hamiltonian are:

|f
Ψ >= D | Ψ > and H̃ = D H D† .

In the simplest case of a single particle state rotated about the z− axis, the transformation is
e 1 , θ1 , φ1 ) = D Ψ(r1 , θ1 , φ1 ) = Ψ(r1 , θ1 , φ1 − α).
Ψ(r

Now consider the transformed Hamiltonian for an infinitesimal change


i ~ i ~ i ~ H] · · ·
H̃ = DHD† = e− h̄ α(n̂·J) He+ h̄ α (n̂·J)
'H− α n̂ · [J,

If the Hamiltonian is invariant under the rotation, then we have H̃ = H, which is the statement of
invariance of the equations of motion under rotations. Consequently,

~ H] = 0 d ~
[J, and J = 0,
dt
which is conservation of angular momentum. This is another example of Emmy Noether’s theorem;
we have an invariance under the rotation group leading to a conservation law.
We see that the wavefunctions, Ψ(t),
e Ψ(t) evolve in time (i.e. for times after the transformation to
LAB)
g according to H̃ and H, respectively. Therefore, if invariance holds H̃ = H, the subsequent
time development in LAB and LAB g will be identical and hence no experimental change will occur,
which is the statement of invariance of experimental results. An example of H̃ 6= H would occur if
the LAB
g were located in a strong electric field that was not present in LAB, which would yield
Stark level shifts in LAB.
g
Another related aspect of the equations of motion is the effect on a system’s spectrum as expressed
by the eigenvalues of
E =< Ψ | H | Ψ >=< Ψ e | H̃ | Ψ
e >= E;e
CHAPTER 9. SYMMETRIES 221

however, using the original Hamiltonian we would get a possibly different spectrum

E 0 =< Ψ
e |H|Ψ
e >6= E,

unless we have invariance. That is, the spectrum remains unchanged only if there is invariance of
the Hamiltonian under the canonical transformation. The unitary transformation does not only
apply to the Hamiltonian. It applies to all other operators. What effect does a rotation or a
translation have on other operators?

9.2.4 Translation and Operators


For the Hamiltonian, the transformed operator was obtained using H̃ = D H D† . If H̃ = H, the
i ~ ~
Hamiltonian is said to be invariant under a translation U = e− h̄ ∆·P , and it follows that [P~ , H] = 0.

In general an operator à = U AU will not be invariant under a translation and we therefore
consider the effect of U on a general operator A. Recall that to be a canonical transformation U
must be applied to all states and all operators. The transformed operator is obtained f for an
infinitesimal translation using
i~ ~
à = U AU † ' A − ∆ · [P , A] · · · ,

~ · ∇A,
or with à = A − ∆ ~ we get
h̄ ~
[P~ , A] = (∇A).
i
~ · ∇A
To see that à = A − ∆ ~ is correct, we can expand as follows
i ~ ~ i ~ ~
~ ' A(~r) − ∆
Ã(~r) = e− h̄ ∆·P Ae+ h̄ ∆·P = A(~r − ∆) ~ · ∇A.
~

Operator Change under Translation


Further insight into the last property concerning how an operator changes under translation is
gained by examining a particular example. The simple example is to take the operator to be
A → ~rop itself. The eigenvalue equation A | Ψ >= a | Ψ >, then becomes

~rop Ψ(r) = ~r0 Ψ(r),

where ~r0 is the eigenvalue and Ψ(r) = δ(~r − ~r0 ) is the (unnormalized) g wave function. After
application of U to the above eigenvalue equation, we obtain the à | g Ψ > = a| gΨ > form

~rop Ψ(r)
e e = ~r0 Ψ(r),
e

where Ψ(r)
e is a new function of the old coordinate ~r; Ψ(r)
e is peaked at LAB.
g It is given by

~ P
e r) = e− h̄i ∆· ~ ~ = δ(~r − ~r0 − ∆).
~
Ψ(~ Ψ(~r) = Ψ(~r − ∆)

Now consider the old operator acting on the new function, that is the A | Ψ >= a0 | Ψ > type
equation,
~rop Ψ(r)
e ~ Ψ(r)
= (~r0 + ∆) e = (e ~ Ψ(r).
~rop + ∆) e

From this we see that


~rop = e ~
~rop + ∆
and
~ = U ~rop U † .
~rop = ~rop − ∆
e

Thus we have
i ~ ~ i ~ ~
~
~rop = e− h̄ ∆·P ~rop e+ h̄ ∆·P = ~rop − ∆.
e

which is equivalent to our basic operator equation [pi , rj ] = h̄i δi,j .


f We now use the abbreviation ∆~ ~
x = ∆.
g You could think of this wave function as a sharply-peaked, normalized wave function centered at ~
r0 .
CHAPTER 9. SYMMETRIES 222

The above rules can be generalized to a general operator using


X
A(~rop ) = an (~rop )n
n

and
U A(~rop )U † = A(e ~
~rop ) = A(~rop − ∆)
~ yields our result
which for small ∆
~ · ∇A(~
Ã(~rop ) ' A(~rop ) − ∆ ~ rop ).

We now turn to rotation symmetries where similar observations will be made.


CHAPTER 9. SYMMETRIES 223

9.2.5 Rotation and Operators


i ~
Application of the rotation operator D = e−α h̄ (n̂·J) to a general operator A yields the following
result for the case of an infinitesimal rotation α → δα
i ~ A]
à = DAD† = A − δα n̂ · [J,

or

~ A] = h̄ A − Ã .
n̂ · [J,
i δα
h
For a scalar operator, the transformed operator is given by

~ r A(~r0 ),
~rop ) = A(~rop − δα(n̂ × ~r0 )) ' A(~rop ) − δα(n̂ × ~r0 ) × ∇
à = A(e 0

as will soon be demonstrated. Using this property, the commutator becomes

~ A] = h̄ 1
n̂ · [J, δα(n̂ × ~r0 ) × ∇r0 A(~r0 ) = n̂ · (~r × p~) A(~r0 ).
i δα
We conclude that a general scalar operator satisfies
~ A] = (~`A),
[J,

~ r.
where ` ≡ ~r × p~, with p~ = h̄i ∇
~ ~
Since we see that [`, A] = (`A), it follows that the spin operator commutes with a scalar field or
that a scalar field contributes only orbital and no spin angular momentum. i
It remains to show why à = A(~rop − δα(n̂ × ~r0 )) is correct. A simple example is helpful, in analogy
to the previous translation case. Namely, we have a system consisting of a single particle located
precisely at ~r0 which is then subject to a rotation about n̂ by an angle δα.


δα 
pp
 pp
pp 
1s


β

 ~r0


Figure 9.3: The active rotation of a particle at ~r0 by a rotation by δα around an axis n̂. The angle
between ~r0 and n̂ is β. The short vector is the displacement into the page by r0 sin βδα.

The displacement due to the rotation is

r0 sin βδα = (n̂ × ~r0 )δα.

Therefore, from the original wave function Ψ = δ(~r − ~r0 ), we obtain the new wave function

e r) = DΨ(~r) = Ψ(~r − ~r0 − δα (n̂ × ~r0 )).


Ψ(~
h Really
a scalar field operator.
i The
above discussion could be applied to a scalar operator of the type A(pop ) with the same result as above
except we find that ~ ~ p is an operator.
rop → h̄i∇
CHAPTER 9. SYMMETRIES 224

As before, consider the three cases of: (1) original eigenvalue equation A | Ψ >= a | Ψ >; (2) the
new equation à | Ψ
e >= a | Ψe >; and (3) the old operator on the new wave function case
0 e
A | Ψ >= a | Ψ > . These are expressed as
e

~rop Ψ(~r) = ~r0 Ψ(~r),

~˜rop Ψ(~
e r) = ~r0 Ψ(~
e r),

and
~rop Ψ(~ e r) = (~˜rop + δα (n̂ × ~r0 ))Ψ(~
e r) = (~r0 + δα (n̂ × ~r0 ))Ψ(~ e r),

where we apply the old operator to Ψ(~


e r), which is the new function of the old coordinates. From
the above we find that
~˜rop = ~rop − δα (n̂ × ~rop ).
Hence X
Ã(~r) = D A D† = n
an D~rop D† = A(~˜r) = A(~rop − δα (n̂ × ~rop )),
n
n
P
where we assumed that we have an expansion A(~r) = n an~rop .

Vector Operator
The previous steps can now be applied to a vector operator. Consider the transformation of the
λth component of a vector operator Aλ , where λ = x, y, z,
i ~ Aλ ].
Ãλ = D Aλ D† = Aλ − δα n̂ · [J,

We can show that under a rotation with infinitesimal δα
~ λ.
Ãλ = Aλ − δα (n̂ × A)

This follows from


< Ψ | Aλ | Ψ >=< Ψ
e | Ãλ | Ψ
e>

and < Ψe | Aλ | Ψ ~ λ =< Ψ


e >= rotated vector = Aλ + δα (n̂ × A) ~ |Ψ
e | Ãλ + δα (n̂ × Ã) e > . We
λ
conclude that
~ λ.
Ãλ = Aλ − δα (n̂ × A)


δα 
pp
 pp
pp 
1s


β


A~


Figure 9.4: The active rotation of a vector operator Aλ by a rotation by δα around an axis n̂. The
~ and n̂ is β. The short vector is the displacement into the page by δα n̂ × A.
angle between A ~

Using this result, we deduce that


i ~ Aλ ] = Aλ − δα (n̂ × A)
~ λ = Aλ − δα (n̂ × A)
~ · λ̂ = Aλ − δα (n̂ · A)
~ × λ̂,
Aλ − δα n̂ · [J,

CHAPTER 9. SYMMETRIES 225

where λ̂ denotes the unit vectors x̂, ŷ, ẑ. Our general result for the commutator of a vector operator
is thus
h̄ ~
[Jµ , Aλ ] = (A × λ̂)µ = h̄i µλν Aν .
i
Another way to write this is:
[Jx , Ay ] = h̄i Az
an cyclic permutations
[Jy , Az ] = h̄ i Ax ,
etc. We will derive selection rules from these commutators when we take the dipole radiation
operator as our vector operator Aλ → erλ .

9.2.6 Angular Momentum Commutators


i ~ ~ and it followed that
The rotation operator D = e− h̄ α(n̂·J) has been applied to a vector operator A

[Jµ , Aν ] = i h̄ µνλ Aλ

or
[Jx , Ay ] = i h̄ Az ,
~ to be J~ itself, we arrive at the angular momentum commutator
etc. Taking A

[Jµ , Jν ] = i h̄ µνλ Jλ

or
[Jx , Jy ] = i h̄ Jz ,
which is a direct generalization of the orbital angular momentum case.
Another proof of this result can be obtained by making two infinitesimal rotations about the x̂ and
ŷ axes, for example. First consider the rotation in the order: (1) first rotate about the x̂ and (2)
then about the ŷ axis, with associated angles δθx , δθy . Note that for the active rotation convention,
positive angles correspond to counter-clockwise rotation directions (see the earlier example of how
we actively moved the LAB to LAB’). The two rotations are therefore described by first Dx , then a
Dy rotation
i i
Dy Dx = e− h̄ δθy Jy e− h̄ δθx Jx
For simplicity apply these rotations to a vector ~a ≡ 1x̂ as shown in the figure.

ẑ ẑ

δθ
y
 ŷ ŷ


-
~b = ~a
δθx ~b


~c

x̂ x̂

Figure 9.5: The active rotation of a vector ~a = x̂ by an angle δθx around an axis x̂, followed by
rotation of a vector ~b by an angle δθy around an axis ŷ. This represents the rotation Dy Dx described
in the text.

One gets
Dx~a = ~b = ~a
and
Dy Dx~a = Dy ~a = ~c = x̂ − ẑ δθy .
CHAPTER 9. SYMMETRIES 226

Now reverse the order of the rotations

Dy~a = ~c = x̂ − ẑδθy

and
Dx Dy~a = Dx (x̂ − ẑδθy ) = x̂ − (Dx ẑ)δθy .
Using Dx ẑ = ẑ − ŷδθx , we have

Dx Dy ~a = x̂ + δθy δθx ŷ − δθy ẑ.

The difference between the two rotation orders is

(Dx Dy − Dy Dx ) ~a = δθy δθx ŷ.

Also
Dx Dy − Dy Dx = −δθx δθy [Jx , Jy ].
This follows from

Dx Dy = (1 − i δθx Jx ) (1 − i δθy Jy ) (9.2)


= 1 − i δθx Jx − i δθy Jy − δθx δθy Jx Jy
Dy Dx = (1 − i δθy Jy ) (1 − i δθx Jx )
= 1 − i δθx Jx − i δθy Jy − δθx δθy Jy Jx
Dx Dy − Dy Dx = −δθx δθy [Jx , Jy ].

The vector δθy δθx ŷ can now be obtained in terms of a rotation about the ẑ axis with angle
δθz = δθy δθx (these angles are unitless). Take

Dz ~a = e−i δθz Jz
~a = ~a + δθz ŷ = ~a − i δθz Jz ~a,

therefore ŷ = −iJz~a, and we have

(Dx Dy − Dy Dx ) ~a = −i δθx δθy Jz ~a = −δθx δθy [Jx , Jy ].

We obtain
[Jx , Jy ] = ih̄ Jz ,
restoring the h̄ factor. Although this was shown here for a special choice of ~a, this result can be
generalized to arbitrary ~a.
Here we see that the above commutator is a direct consequence of the non-commutativity of
rotations in three dimensions. This property holds in classical mechanics also. The question arises:
where does the quantum mechanics enter? We introduced h̄, but that only established units for
angular momentum. The answer is: quantum mechanics enters when we associate J~ with the
angular momentum operator and use it to obtain eigenvalue equations and hence quantize values
of angular momentum in multiples of h̄.

9.2.7 Addition of Angular Momentum


Having deduced the commutators for angular momentum, including orbital and spin motion, we
can now discuss the problem of adding angular momentum from a general viewpoint. Many of our
operator steps were given before when we dealt with the orbital angular momentum case, so I will
simply review that material.

Review–Omitted from lectures


Define J 2 = J~ · J~ = Jx2 + Jy2 + Jz2 and J± = Jx ± iJy . It follows from [Jµ , Jν ] = i h̄µνλ Jλ , that
[J 2 , Jz ] = 0 ; therefore, we can measure J 2 and Jz simultaneously with infinite precision. They are
commuting observables which satisfy eigenvalue equations

J 2 | j m >= h̄2 j(j + 1) | j m >


CHAPTER 9. SYMMETRIES 227

and
Jz | j m >= h̄m | j m >,
where j and m denote the eigenvalues. From these eigenvalue equations for Hermitian operators, it
follows that < j 0 m0 | j m >= δj 0 j δm0 m . The states | j m > provide a basis for which both J 2 and
Jz are diagonal.

< j 0 m0 | J 2 | j m > = = h̄2 j(j + 1) δj 0 j δm0 m (9.3)


0 0
< j m | Jz | j m > = = h̄m δj 0 j δm0 m .
(9.4)

We also have [J± , Jz ] = ∓J± h̄ and [J± , J 2 ] = 0 as a consequence of the basic commutator. j
These commutators can be used to show that J± are raising and lowering operators for angular
momentum projection m, but do not change the value of j. One has

J 2 J± | j m >= J± J 2 | j m >= h̄2 j(j + 1) J± | j m >,

therefore J± | j m > is an eigenstate of J 2 . Consider a state | jm0 >, with a particular value for
m = m0 then since

Jz J− | j m0 > = (Jz J− − J− Jz + J− Jz ) | j m0 > (9.7)


= h̄m0 J− | j m0 > −[J− , Jz ] | j m0 >
= h̄(m0 − 1)J− | j m0 > .

Therefore J− | j m0 > is an eigenstate of Jz with eigenvalue (m0 − 1), and we can write

J− | j m0 >= α | j m0 − 1 >,

where a parameter α is included as a phase and for normalization. Note that J− | j m0 > is still
an eigenstate of J 2 with eigenvalue j(j + 1). Similar steps with J+ lead to

J+ | j m0 >= β | j m0 + 1 >,

where a parameter β is included as a phase and for normalization.


The values of α and β are obtained from

< j m | J− J+ | j m >=| β |2 < j m + 1 | j m + 1 >=| β |2 = [j(j + 1) − m(m − 1)]h̄2

and

< j m | J+ J− | j m >=| α |2 < j m − 1 | j m − 1 >=| α |2 = [j(j + 1) − m(m + 1)]h̄2


j The following relations are often useful
[J± , Jz ] = [Jx ± Jy , Jz ] (9.5)
= −h̄iJy ± ih̄ i, Jx = ∓(Jx ± iJy )
= ∓ i J± .

J2 = Jx2 + Jy2 + Jz2 (9.6)


= Jz2 + (Jx − i Jy ) (Jx + i Jy ) + Jz h̄
J− J+ = Jx2 + Jy2 + i[Jx , Jy ] = (J− J+ )†
J+ J− = Jx2 + Jy2 − i[Jx , Jy ] = (J+ J− )†

J− = J+ .

Also
[J− , J+ ] = 2i[Jx , Jy ] = −2h̄Jz ,
and
1
J 2 = Jz (Jz + h̄) + J− J+ = Jz (Jz − h̄) + J+ J− = Jz2 + (J− J+ + J+ J− ).
2
CHAPTER 9. SYMMETRIES 228

Therefore, p
J± | j m >= 1h̄ j(j + 1) − m(m ± 1) | j m ± 1 >,
where the phase choice of Condon & Shortley is used, which is the usual choice made to generate
simple real coefficients for the addition of angular momentum.
The raising and lowering process allows us to obtain a set of eigenstates by stepping up and down
in units of h̄. The question arises: can this stepping process continue indefinitely? The answer is:
No! It is limited by the finite value of the angular momentum < j m | J 2 | j m >= h̄2 j(j + 1).
..
.
| j m0 + 2 >
| j m0 + 1 >
| j m0 >
| j m0 − 1 >
| j m0 − 2 >
..
.

Figure 9.6: The levels generated by the raising J+ and lowering J− operators acting on an initial
state | j m0 > .
k
The explicit bounds on the m values follow from
< j m | J− J+ | j m >=< j m | J 2 − Jz (Jz + h̄) | j m >= h̄2 [ j(j + 1) − m(m + 1) ] ≥ 0.
Therefore, m(m + 1) ≤ j(j + 1) places a bound on m. Suppose we have the largest value of
m = m> , then
J 2 | j m> >= j(j + 1)h̄2 | j m> >= h̄2 m> (m> + 1) | j m> > +J− J+ | j m> > .
Since J+ | j m> >= 0, we have
j(j + 1) = m> (m> + 1).
Similar steps applied to the lowest value of m = m< , yield
j(j + 1) = m< (m< − 1)
and now
m> (m> + 1) = m< (m< − 1)
or
(m> + m< )(m> − m< + 1) = 0
m> = m< − 1 or m> = −m< .
Only the second choice is viable.
We can start with | jm> > and apply (J− )n n times to reach | jm< > in n steps. Therefore
m> − n = m< . or
(m> − m< ) = n = 2 m>
and
n
= j.
m< =
2
Since j(j + 1) = m> (m> + 1), we have j = m> . The allowed values of j are therefore
0, 21 , 1, 32 , 2, 52 , 4, · · · .
In this PNreview, the angular moment J was not limited to one particle. Since we obtained
J~ = n=1 (~`n + ~sn ), earlier and we obtained the commutators for this total angular momentum,
the above results apply to any system of n particles, including all spins.
The question arises: How can we construct eigenstates | JM > . for n particles in terms of the
basic angular momenta. Also how can we add the angular momenta of two complicated systems?
k Some useful properties here are: J 2 = Jz (Jz + h̄) + J− J+ = Jz (Jz − h̄) + J+ J− . and [J+ , J− ] = 2h̄Jz .
CHAPTER 9. SYMMETRIES 229

Coupling of Two Angular Momenta


Consider two systems and their corresponding angular momenta operators ~j(1) and ~j(2). These
systems could refer to elementary particles or systems with internal structure, such as atoms. l
Here ~j refers to the the orbital, spin, or total angular momenta of the system. It is now assumed
that ~j(1) and ~j(2) are separate, unchanging physical attributes of each system, so they commute

[~j(1), ~j(2)] = 0.

If this property were not valid, our rules for adding angular momenta would have to be modified or
simply applied to subsystems for which the above rule does hold.
The sum of the angular momenta is defined by J~ = ~j(1) + ~j(2), which satisfies

[Jx , Jy ] = i h̄Jz = [jx (1) + jx (2), jy (1) + jy (2)] = i h̄(jz (1) + jz (2)).

Note how important [~j(1), ~j(2)] = 0 is in this derivation.


All the other relations now follow:

[J 2 , Jz ] = 0; [J± , Jz ] = ∓J± h̄; [J± , J 2 ] = 0

and
J 2 = Jz (Jz − h̄) + J+ J− = Jz (Jz + h̄) + J− J+
p
J± | J M >= h̄ J(J + 1) − M (M ± 1) | J M ± 1 >
We can now form states that have definite J 2 and Jz eigenvalues

J 2 | J M >= h̄2 J(J + 1) | J M >

Jz | J M >= h̄M | J M > .


The problem is to use the | j1 m1 > and | j2 m2 > to construct the basis states | J M > .
The product basis states give us the “product representation” for the two spin states

| j1 m1 j2 m2 >≡| j1 m1 > | j2 m2 >,

which are eigenstates of the commuting (mutually compatible) observables j12 , j1z , j22 , j2z . There
are (2j1 + 1) ⊗ (2j2 + 1) product states. Our job is to find linear combinations of these product
representation states to form the “irreducible representation,” i,e, the eigenstates of j12 , j22 , J 2 and
Jz . m This represents a change in basis or a unitary transformation from | j1 m1 j2 m2 > to
| j1 j2 JM >≡| JM > .
To start this project, let us note that the product basis states are already eigenstates of Jz since

Jz | j1 m1 j2 m2 >= [jz (1) + jz (2)] | j1 m1 >| j2 m2 >= h̄(m1 + m2 ) | j1 m1 j2 m2 > .

However, because of the ~j1 · ~j2 term in J 2 ,

J 2 = j12 + j22 + 2 ~j1 · ~j2 = j12 + j22 + 2jz (1) jz (2) + j− (1)j+ (2) + j+ (1)j− (2),

the product states are not eigenstates of J 2 . Various values of m1 and m2 get mixed in by
j+ (1)j− (2) operators. The operators j12 and j22 are already “diagonal,” which will not be altered by
the following discussion.
The linear combinations are needed to generate eigenstates of J 2 . The coefficients of this linear
combination of product states are called:
• Clebsch-Gordan coefficients (CG)
• Wigner coefficients (3-j symbols)
• Vector Addition coefficients.
l Note that some so called “Elementary” particles, such as nucleons, also have structure and excitations.
m Irreducible means | JM > transforms only to other states with the same J value under rotations; no other J−
values enter in D | JM >=| JM >0 .
CHAPTER 9. SYMMETRIES 230

Clebsch-Gordan Coefficients
The product states are combined linearly to construct eigenstates of the four quantum numbers
J 2 , Jz , j12 , j22 out of the eigenstates of the four quantum numbers j12 , jz (1), j22 , jz (2); we write
X X
| j1 j2 JM >= | j1 m1 j2 m2 >< j1 m1 j2 m2 | j1 j2 JM >= < j1 m1 j2 m2 | JM >| j1 m1 j2 m2 >,
m1 ,m2 m1 ,m2

where we set < j1 m1 j2 m2 | j1 j2 JM >→< j1 m1 j2 m2 | JM > in the oft-used notation for the
Clebsch-Gordan coefficient < j1 m1 j2 m2 | JM > . We see that |< j1 m1 j2 m2 | JM >|2 is the
probability for finding the particles with m1 , m2 values in the state | j1 j2 JM > with definite total
angular momentum and projection J, M.
Several properties of these Clebsch-Gordan coefficients are already known or can be deduced:

1. Since Jz is already diagonal in the product representation, we have < j1 m1 j2 m2 | JM >= 0,


unless M = m1 + m2 .

2. The maximum value of M and hence of J is J0 = max(M ) = j1 + j2 , obtained with m1 = j1


and m2 = j2 . Hence,
| j1 j2 J0 J0 >= + | j1 j1 > | j2 j2 > .
with the overall plus sign chosen as a convention.
3. The possible values of J are

j1 + j2 , j1 + j2 − 1, j1 + j2 − 2, · · · | j1 − j2 |,

which is denoted by the triangle condition ∆(j1 , j2 , J) = 0. See later for proof of this rule.
4. The total number of states are the same in the two representations. From the previous range
of J values as given by the ∆(j1 , j2 , J) = 0 condition, we have the sum
jX
1 +j2

(2J + 1) = (2j1 + 1)(2j2 + 1).


J=|j1 −j2 |

Hence the total number of states in the two representations is the same.
5. The transformation from the “product” to the “irreducible” representation is a unitary
transformation, since it maps one complete orthonormal basis to another. Hence, we can
write not only
X
| j1 j2 JM >= < j1 m1 j2 m2 | JM >| j1 m1 j2 m2 >,
m1 ,m2

but also the inverse transformation


X
| j1 m1 j2 m2 >= < JM | j1 m1 j2 m2 >| j1 j2 JM > .
J,M

It follows that the unitary conditions are (U † U = 1) or


X
< JM | j1 m1 j2 m2 > < j1 m1 j2 m2 | J 0 M 0 >= δJJ 0 δM JM 0 ,
m1 ,m2

and (U U † = 1) or
X
< j1 m1 j2 m2 | JM > < JM | j1 m01 j2 m02 > = δm1 m01 δm2 m02 ,
J,M
CHAPTER 9. SYMMETRIES 231

6. The phase conventions usually adopted n permit us to have real CG coefficients


< j1 m1 j2 m2 | JM >∗ =< j1 m1 j2 m2 | JM > . Furthermore, we have < a | b >=< b | a >∗ for
this case since X
| j1 j2 JM >= < j1 m1 j2 m2 | JM >| j1 m1 j2 m2 >
m1 ,m2
X X
< j1 j2 JM |= < j1 m1 j2 m2 | JM >∗ < j1 m1 j2 m2 |= < JM | j1 m1 j2 m2 >< j1 m1 j2 m2 | .
m1 ,m2 m1 ,m2

It follows that < j1 m1 j2 m2 | JM >=< j1 m1 j2 m2 | JM >∗ =< JM | j1 m1 j2 m2 > .


7. The CG coefficients can be determined either from recursion relations, o a closed form
formula of Racah, or by looking them up in standard tables. Some standard resources are
Messiah (Apppendix), Brink & Satchler, or using Mathematica.

The three-j (3-j) or Wigner coefficients are defined in a more symmetric way than are the CG
coefficients.

 
j1 −j2 +M j1 j2 J
< j1 m1 j2 m2 | JM >≡ 2J + 1(−1) .
m1 m2 −M
Symmetry properties of the 3-j’s are simply given as:
     
j1 j2 j3 j2 j3 j1 j2 j1 j3
= = (−1)j1 +j2 +j3 .
m1 m2 m3 m2 m3 m1 m2 m1 m3

The 3-j’s are also tabulated in the above references. These symmetries imply corresponding
symmetries in the CG coefficients.

Proof of Triangle Rule


Apply J− to | j1 j2 J0 J0 > where J0 = j1 + j2 .
p
J− | j1 j2 J0 J0 >= J0 (J0 + 1) − J0 (J0 − 1) | j1 j2 J0 J0 − 1 >

= [j− (1) + j− (2)] | j1 j1 >| j2 j2 >


p p
= 2j1 | j1 j1 − 1 >| j2 j2 > + 2j2 | j1 j1 >| j2 j2 − 1 > .
Therefore we get the normalized state
r r
j1 j2
| j1 j2 J0 J0 − 1 >= | j1 j1 − 1 >| j2 j2 > + | j1 j1 >| j2 j2 − 1 > .
J0 J0

We have lowered M but not J0 ; the above state is still an eigenstate of J 2 with J0 = j1 + j2 , since
[J 2 , J− ] = 0. We can continue to apply J− and obtain the 2J + 1 = 2(j1 + j2 ) + 1 states with
J0 = j1 + j2 .
Now the following combination can be made to form the state with J = j1 + j2 − 1 = J0 − 1 = M :
r r
j2 j1
| j1 j2 J0 − 1J0 − 1 >= − | j1 j1 − 1 >| j2 j2 > + | j1 j1 >| j2 j2 − 1 > .
J0 J0
This conclusion can be verified by noting the orthogonality

< j1 j2 J0 J0 − 1 | j1 j2 J0 − 1J0 − 1 >= 0


n The phase conventions used consist of (1) taking the Condon-Shortley phase in J | JM >=
p ±
+1 J(J + 1) − M (M ± 1); using | Jmax Jmax >= +1 | j1 j1 >| j2 j2 >; and (3) taking < j1 j1 j2 J − j1 | JJ > > 0,
and real. Once these three conventions are used the other CG coefficients are made real by virtue of the recursion
relations.
o See H.W. for a proof of
p p
j1 (j1 + 1) − m1 (m1 ∓ 1) < j1 m1 ∓ 1j2 m2 | JM > + j2 (j2 + 1) − m2 (m2 ∓ 1) < j1 m1 j2 m2 ∓ 1 | JM >=
p
J(J + 1) − M (M ± 1) < j1 m1 j2 m2 | JM ± 1 >, M ± 1 = m1 + m2 .
CHAPTER 9. SYMMETRIES 232

and that
J+ | j1 j2 J0 − 1J0 − 1 >= 0,
which states that the maximum M value is J0 − 1 for this state. Note the phase convention
< j1 j1 j2 J − j1 | JM > > 0, has been used here.
The stepping down process can be continued. Applying J− to both of the above states yields the
J = J0 and J = J0 − 1 eigenstates of one lower M = J0 − 2 | j1 j2 J0 J0 − 2 > and
| j1 j2 J0 − 1J0 − 2 > . These two states are linear combinations of the product states
| j1 j1 − 2 >| j2 j2 >; | j1 j1 − 1 >| j2 j2 − 1 >; | j1 j1 >| j2 j2 − 2 > . A third state orthogonal to the
J = J0 and J = J0 − 1 states can be formed from the above three product states. This new state
has J = J0 − 2 = M, | j1 j2 J0 − 2, J0 − 2 > . Continuation of this process leads to J values of
J = j1 + j2 , j1 + j2 − 1, j1 + j2 − 2 · · · | j1 − j2 |, which corresponds to the vector addition rules of
Lande-Sommerfeld. We denote this result by ∆(j1 , j2 , J) = 0, which means the CG vanished unless
this addition rule is satisfied. A more general proof will follow later.

9.2.8 Addition of Three Angular momenta ( Racah Coefficients or 6-j


Symbols)
Let us briefly consider the addition of three distinct angular momenta, which will lead us to the
Racah Coefficients or 6-j Symbols, which often enter into the analysis of finite systems with
angular momenta quantum numbers. The basic idea is that three distinct angular momenta can be
added together in different ways–using various coupling schemes. One way is

J~ = ~j1 + ~j2 + ~j3 = J~12 + ~j3 ,

or
X
| (j1 j2 )J12 j3 ; JM >= < J12 , M12 j3 m3 | JM >< j1 m1 j2 m2 | J12 M12 >| j1 m1 > | j2 m2 > | j3 m3 > .
m1 ,m2 ,m3 ,M12

Another way is, for example,


J~ = ~j1 + ~j2 + ~j3 = ~j1 + J~23 ,
or
X
| j1 (j2 j3 )J23 ; JM >= < j1 m1 J23 , M23 | JM >< j2 m2 j3 m3 | J23 M23 >| j1 m1 > | j2 m2 > | j3 m3 > .
m1 ,m2 ,m3 ,M23

For each state there are 6 “good” quantum numbers, i.e. six operators that satisfy eigenvalue
equations and are “ diagonal.” They are: for the product representation (j12 j1z j22 j2z j32 j3z ); for the
first coupling scheme (j12 j22 J12 j32 J Jz ); and for the second coupling scheme (j12 j22 j32 J23 J Jz ).
The overlap between the two coupling schemes, or the transformation of one coupling scheme into
another, is given by the Racah coefficient (W) or equivalently by the 6-j symbol defined by
X
< j1 (j2 j3 )J23 ; J 0 M 0 | (j1 j2 )J12 j3 ; JM >= δJJ 0 δM M 0
m1 ,m2 ,m3 ,M12 ,M23

< JM | j1 m1 J23 M23 >< J23 M23 | j2 m2 j3 m3 >< j1 m1 j2 m2 | J12 M12 >< J12 M12 j3 m3 | JM >
1 1
= (2J12 + 1) 2 (2J23 + 1) 2 W (j1 j2 Jj3 ; J12 J23 )
1 1  
(2J12 + 1) 2 (2J23 + 1) 2 j1 j2 J12
=
(−1)j1 +j2 +j3 +J j3 J J23
The 6-j symbol is defined to give a simple form for symmetries under exchange of various angular
momenta. The above overlap can be shown ( using J+ and J 2 operators to be independent of the
M − value and to be nonzero only if J = J 0 and M = M 0 .
CHAPTER 9. SYMMETRIES 233

9.2.9 Addition of Four Angular momenta ( 9-j Symbols)


Extending the above ideas to the addition of four distinct angular momenta leads us to the 9-j
Symbols, which are related to the transformation of one coupling scheme to another. The basic
idea is that four distinct angular momenta can be added together in different ways–using various
coupling schemes.
One way is
J~ = ~j1 + ~j2 + ~j3 + ~j4 = J~12 + J~34 ,
or X
| (j1 j2 )J12 (j3 j4 )J34 ; JM >= < J12 , M12 J34 , M34 | JM >
m1 ,m2 ,m3 ,m4 ,M12 ,M34

< j1 m1 j2 m2 | J12 M12 >< j3 m3 j4 m4 | J34 M34 >| j1 m1 > | j2 m2 > | j3 m3 > | j4 m4 > .
Another coupling is
J~ = ~j1 + ~j2 + ~j3 + ~j4 = J~13 + J~24 ,
or X
| (j1 j3 )J13 (j2 j4 )J24 ; JM >= < J13 , M13 J24 , M24 | JM >
m1 ,m2 ,m3 ,m4 ,M13 ,M24

< j1 m1 j3 m3 | J13 M13 >< j2 m2 j4 m4 | J24 M24 >| j1 m1 > | j2 m2 > | j3 m3 > | j4 m4 > .

*A@
 *@
J~12  J~ J~13  J~
 A
 ~j2 A@ 34  ~j3 A@ 24
~1: ~j3A @ ~ ~j2A @
j  AA
UP ~j4@ 
j

1:
 AAU ~j4@
 PP@   PP@

  qR
@
- 
  q-
PR
@
J~ J~
Figure 9.7: Addition of four spins in two coupling schemes.

For each state there are 8 “good” quantum numbers, i.e. eight operators that satisfy eigenvalue
equations and are “ diagonal.” They are: for the product representation (j12 j1z j22 j2z j32 j3z j42 j4z ); for
the first coupling scheme (j12 j22 j32 j42 J12 J34 JJz ); and for the second coupling scheme
(j12 j22 j32 j42 J13 J24 JJz ).
Changing from one coupling scheme to another is given by the overlap or 9-j symbol defined by

X
< (j1 j3 )J13 (j2 j4 )J24 ; J 0 M 0 | (j1 j2 )J12 (j3 j4 )J34 ; JM >= δJJ 0 δM M 0
m1 ,m2 ,m3 ,m4 ,M12 ,M34 ,M13 ,M24

< J13 , M13 J24 , M24 | JM >< j1 m1 j3 m3 | J13 M13 >< j2 m2 j4 m4 | J24 M24 >
< J12 , M12 J34 , M34 | JM >< j1 m1 j2 m2 | J12 M12 >< j3 m3 j4 m4 | J34 M34 >
1 1
= (2J12 + 1) 2 (2J23 + 1) 2 W (j1 j2 Jj3 ; J12 J23 )
 
1
 j1 j2 J12 
= [(2J12 + 1)(2J34 + 1)(2J13 + 1)(2J24 + 1)] 2 j3 j4 J34
J13 J24 J
 

The 9-j symbol is defined to give a simple form for symmetries under exchange of various angular
momenta. It is a simple way to store selection rules as we shall see later when we consider matrix
elements describing physical processes.
CHAPTER 9. SYMMETRIES 234

An important example of the change in coupling scheme is the transformation from the j-j coupling
scheme (used for the case where there is a strong spin-orbit coupling) to the S-L (
Saunders-Russell) coupling scheme where the spin orbit is relatively weak. Then we have
1
X
| (sa sb )S, (`a `b )L; JM >= [(2S + 1)(2L + 1)(2ja + 1)(2jb + 1)] 2
ja jb
 
 sa sb S 
`a `b L | (sa `a )ja , (sb , `b )jb ; JM > .
ja jb J
 

9.2.10 Matrix Representation of the Rotation Group


The state vector for a system of particles has been expressed as eigenstates of the total angular
momentum operators J 2 and Jz . These states | JM > have been constructed by forming linear
combinations of the product states and are now used as a set of basis states for expressing the
i
unitary rotation operator D = e− h̄ α(n̂·J) in a matrix representation. For a general operator A and
a basis state | n >, the matrix representation is given by
X
A= | n >< n | A | n0 >< n0 |,
nn0

where A is an operator in the Hilbert space and < n | A | n0 > is a matrix representation of that
operator.
For the rotation operator, using the basis states | JM >,we obtain
X
D= | JM >< JM | D | J 0 M 0 >< J 0 M 0 |,
JM,J 0 M 0

X
J 0
= | JM > DM M 0 (α, β, γ) < JM |, .
JM M 0

~ = 0 to show that only J 0 = J matrix elements occur and we


Here we use the commutator [J 2 , J]
have defined the Wigner D function as

< JM | D(α, β, γ) | J 0 M 0 >≡ DM


J
M 0 (α, β, γ).

The three angles (α, β, γ) are the Euler angles that we will introduce next to replace the original
i
three angles in D = e− h̄ α(n̂·J) . The Euler angles are needed to describe any finite rotation
conveniently.
J
We see that the matrix DM M 0 breaks into blocks corresponding to a given value of J, no further
reduction can be formed which do not mix under a rotation. Therefore we refer to this breakdown
of D as a irreducible representation and the corresponding states are called an irreducible basis.
The block form is:  
D J1 0 0 0
DJ2
 
 0 0 0 
 .
 0
 0 DJ3 0  
..
0 0 0 .
For spin 1/2 added to spin 1/2, the reduction process is represented by a 4 × 4 matrix reduced to a
3 × 3 plus a 1 × 1 matrix with S = 1 and S = 0 respectively.

< 1/2m1 1/2m2 | D | 1/2m01 1/2m02 >→< SMS | D | SMs0 S >

which is of the form    


x x x x y y y 0
 x x x x   y y y 0 
 → 
 x x x x   y y y 0 
x x x x 0 0 0 z
CHAPTER 9. SYMMETRIES 235

where x are the entries for the product representation matrix and y and z are the triple S = 1 and
singlet S = 0 irreducible representation entries.
This process of reduction is symbolized by the expression
X
Dj1 ⊗ Dj2 = DJ .
J

9.2.11 Euler Angles and Finite Rotations


References:
• Brink, D. M. & Satchler, G. R., “Angular Momentum.”
• Also, for more than you want to know about Euler angles: Eric W. Weisstein.“Euler Angles.”
From MathWorld–A Wolfram Web Resource.
http://mathworld.wolfram.com/EulerAngles.html
We wish to express a given finite rotation of the system (an active rotation) in terms of Euler
angles. Recall that
0
~
e >= D | Ψ >= e−i αh̄ (n̂·J)
|Ψ ,
symbolizes this rotation, where three angles ( α0 and the two angles for the axis n̂ ) stipulate a
given rotation. It is convenient to use Euler angles ( for reasons that will become clearer later)
instead of α0 and n̂.
Euler angles are defined by the following steps:
step 1 Originally both space fixed and body-fixed axis are lined up. Now rotate the system about
the body fixed axis ẑ by an angle α. We have: (xyz) → (x1 y1 z1 ≡ z). Note that α is the
α
azimuthal angle φ. This rotation is Dz (α) = e−i h̄ Jz .

step 2 Next rotate the system about the body fixed axis y1 by an angle β. We have:
β
(x1 y1 z) → (x2 y1 z2 ). Note that β is the polar angle θ. This rotation is Dy1 (β) = e−i h̄ Jy1 .

step 3 Next rotate the system about the body fixed axis z2 by an angle γ. We have:
γ
(x2 y1 z2 ) → (x3 y3 z3 ≡ z2 ). This rotation is Dz2 (γ) = e−i h̄ Jz2 .

The full rotation operator is thus


γ β α
D(α, β, γ) = e−i h̄ Jz2 e−i h̄ Jy1 e−i h̄ Jz ,

or
D(α, β, γ) = Dz2 (γ)Dy1 (β)Dz (α),
all in terms of body-fixed axes. We label the rotation by R = α, β, γ, where the steps are ordered
in the → direction. On the left hand side the corresponding operators act in the ← sequence.
Note that α → φ, is the azimuthal angle of z2 with respect to the space-fixed xy− axes;β → θ, is
the polar angle of z2 with respect to the space-fixed z− axis.
The inverse rotation, denoted by the sequence R−1 = −γ, −β, −α, returns the rotated frame to the
original xyz location, with
α β γ
D(R−1 ) ≡ D(−γ, −β, −α) = ei h̄ Jz ei h̄ Jy1 ei h̄ Jz2 ,

or
D(α, β, γ) = Dz (−α)Dy1 (−β)Dz2 (−γ) = D−1 (α, β, γ) =
D† (α, β, γ) = D−1 (R) = D† (R).
In a sequence of rotations, first R1 = {α1 , β1 , γ1 } followed by R2 = {α2 , β2 , γ2 }, one must set up a
new axis coinciding with the original space-fixed axis after each sequence. Also, note that the final
orientation after the sequence R1 the R2 could be obtained by one rotation R3 = {α3 , β3 , γ3 },
which is a statement of the group property.
CHAPTER 9. SYMMETRIES 236

STEP 1 STEP 2
Z=Z1 Z1
Z2
D E
Ey
y
D
Y Y
D x x
X X E
x
STEP 3 RESULT
Z1 Z
Z3 Z3

y y
J y
Y Y

X X
J
x x x

Figure 9.8: The Euler angle definition using body-fixed axes of rotation. In step one, the system is
rotated about Z by and angle α, so that X, Y, Z → X1 , Y1 , Z1 = Z. In step two, the system is rotated
about the body-fixed axis y1 by angle β, so that X1 , Y1 , Z1 → X2 , Y2 = Y1 , Z2 . Finally, in step three
the system is rotated about the body-fixed axis Z2 by an angle γ, so that X2 , Y2 , Z2 → X3 , Y3 , Z3 =
Z2 . The net result in shown in the “RESULTS” plot as going from X, Y, Z to X3 , Y3 , Z3 .

Space-fixed Axes
It is convenient to use space-fixed axes to define rotations instead of the result

D(α, β, γ) = Dz2 (γ)Dy1 (β)Dz (α),

found in defining Euler angles rotations about body-fixed axes. Fortunately, it is possible to arrive
at the orientation (x3 , y3 , z3 ) defined by the three Euler angles (α, β, γ) using only rotations about
space-fixed (x, y, z) axes.
Indeed, we shall prove that the above D(α, β, γ) can also be represented by
α β γ
D(α, β, γ) = Dz (α)Dy (β)Dz (γ) = e−i h̄ Jz e−i h̄ Jy e−i h̄ Jz ,

about space-fixed axes. This extremely important theorem can be proven algebraically, but a
geometric illustration is also attached to make it more convincing.
First the algebraic proof is obtained by recalling that a transformed operator à is obtained, after a
unitary transformation is performed, using

à = DAD† ,
CHAPTER 9. SYMMETRIES 237

which gives the operator to be used after a rotation D. Therefore, we can represent the rotation
about y1 in step 2 as a “transformed” operator in the following way
β α β α
Dy1 (β) = e−i h̄ Jy1 = e−i h̄ Jz e−i h̄ Jy e+i h̄ Jz = Dz (α) Dy (β) Dz† (α).

Similarly, the rotation about z2 by γ in step 3 is obtained after two previous unitary
transformations and is hence given by
γ
Dz2 (γ) = e−i h̄ Jz2 = [Dy1 (β) Dz (α) ]Dz (γ) [Dy1 (β) Dz (α) ]† ,
γ
with Dz (γ) = e−i h̄ Jz .
Using these relations in our original form for D(α, β, γ), we get

D(R) = D(α, β, γ) = Dz2 (γ) Dy1 (β) Dz (α) (9.8)


= [Dy1 (β) Dz (α) ]Dz (γ) [Dy1 (β) Dz (α) ]† Dy1 (β) Dz (α)
= Dy1 (β) Dz (α) Dz (γ) Dz† (α) Dy†1 (β) Dy1 (β) Dz (α)
= Dy1 (β) Dz (α) Dz (γ) Dz† (α) Dz (α)
= Dy1 (β) Dz (α) Dz (γ)
= Dz (α) Dy (β) Dz† (α) Dz (α) Dz (γ)
= Dz (α) Dy (β) Dz (γ),

where the unitarity property D† D = 1, has been used several times. The final line expresses the
rotation about the space-fixed z and y axes only.

Geometric Illustrations

A geometric illustration of this result serves as a check. It is rather difficult to visualize why the
space fixed rotations given above are in general equivalent to the original D obtained directly from
the body fixed definitions of the Euler angle rotations.
Some special cases are clearly true; for example, for β = 0 z2 = z and

D(α, β = 0, γ) = Dz2 (γ) Dz (α) = Dz (γ) Dz (α) = Dz (α + γ)

which is exactly the same as the space-fixed expression

D(α, β = 0, γ) = Dz (α) Dy (β = 0) Dz (γ) = Dz (α) Dz (γ).

For α = 0, wherein y1 = y we must show that D(α = 0, β, γ) = Dz2 (γ) Dy (β) is equivalent to
Dy (β) Dz (γ), which is illustrated by the figure.

9.2.12 Wigner Rotation Matrix


We have found that the space fixed rotations

D(α, β, γ) = Dz (α) Dy (β) Dz (γ),

transforms a system from (x, y, z) → (x0 , y 0 , z 0 ) using Euler angles α, β, γ. p


The Wigner operator
D, which now refers to space-fixed axes is still unitary since

D† (α, β, γ) D(α, β, γ) = Dz† (γ) Dy† (β) Dz† (α) Dz (α) Dy (β) Dz (γ) (9.9)
= Dz† (γ) Dy† (β) Dy (β) Dz (γ)
= Dz† (γ) Dz (γ) = 1.

We also have D(α, β, γ) D† (α, β, γ) = 1. The inverse rotation R−1 = (−γ, −β, −α), tells us that

D−1 (α, β, γ) = D(−γ, −β, −α) = D† (α, β, γ),


p We now set (x3 , y3 , z3 ) → (x0 , y 0 , z 0 ).
CHAPTER 9. SYMMETRIES 238

Figure 9.9: For α = 0 rotations about body-fixed axes, Dz2 (γ) Dy (β), and space-fixed axes,
Dy (β) Dz (γ), are seen to be equivalent.

or D−1 (R) = D(R−1 ).


We now define the Wigner rotation matrix by
< JM | D(α, β, γ) | J 0 M 0 >≡ DM
J
M 0 (α, β, γ),

which is a matrix representation of the rotation group in 3-D (R3 ). Since [J 2 , D] = 0, only J 0 = J
matrix elements are non zero. We can now write the various properties of this matrix which follow
CHAPTER 9. SYMMETRIES 239

from the properties of the D− operators: from


< JM | D† (α, β, γ) | J 0 M 0 >=< JM 0 | D(α, β, γ) | JM 0 >∗
=< JM | D−1 (α, β, γ) | J 0 M 0 >=< JM | D(−γ, −β, −α) | J 0 M 0 >,
we obtain
J∗ J
DM 0 M (α, β, γ) = DM M 0 (−γ, −β, −α),

J∗ J −1
or DM 0 M (R) = DM M 0 (R ).
The unitary property DD† = 1, yields
X
J J∗
DM M 00 (R) DM 0 M 00 (R) = δM M 0 .
M 00

The unitary property D D = 1, yields
X
J∗ J
DM 00 M (R) DM 00 M 0 (R) = δM M 0 .

M 00

A property that is characteristic of groups is the sequence of rotations D(R2 ) D(R1 ) being a
member of the group, i.e.
D(R2 ) D(R1 ) = D(R3 ),
or
D(α2 , β2 , γ2 ) D(α1 , β1 , γ1 ) = D(α3 , β3 , γ3 ),
we say that R3 = R2 R1 . This group property is now
X
J J J
DM M 00 (R2 ) DM 00 M 0 (R1 ) = DM M 0 (R3 ).
M 00
J
Note that DM 00 M 0 (0, 0, 0) = 1, which is one required property to define a group.

Other important properties of the D functions are:


J
DM M 0 (α, β, γ) = < JM | Dz (α) Dy (β) Dz (γ) | JM 0 > (9.10)
β γ
−i α
h̄ Jz −i h̄ Jy −i h̄ Jz 0
= < JM | e e e | JM >
0 β
= e−i(αM +γM )
< JM | e −i h̄ Jy
| JM 0 >
0
= e−i(αM +γM )
dJM M 0 (β),
where β
dJM M 0 (β) ≡< JM | e−i h̄ Jy | JM 0 >
is a real orthogonal matrix.
The orthogonality of dJM M 0 follows from
iγ(M −M 0 )
X X
J∗ J
DM 00 M (R) DM 00 M 0 (R) = δM M 0 = e dJM 00 M (β)dJM 00 M 0 (β),
M 00 M 00

or simply
X
dJM 00 M (β)dJM 00 M 0 (β) = δM M 0 .
M 00

The reality of dJM 00 M (β)


is deduced from
p the reality of iJy which in turn occurs because of the
relation < JM | (Jx ± iJy | JM 0 >= J(J + 1) − M (M ± 11) δM ±1,M 0 .
Other important properties ( see Brink & Satchler p122 ff) of the “little” Wigner function
dJM 00 M (β) are:

0
dJM M 0 (β) = (−1)M −M dJM 0 M (β) = dJ−M 0 −M (β) (9.11)
= dJM 0 M (−β) = (−1)J+M dJM −M 0 (π − β)
0
= (−1)J+M dJM −M 0 (π + β).
These symmetries play an important role in allowed states and selection rules for composite
systems.
CHAPTER 9. SYMMETRIES 240

9.2.13 Spherical Harmonics and The Wigner Rotation Matrix


To make the Wigner rotation functions more familiar, let us consider the spherical harmonic
Y`m (θφ) as the angular part of the wave function of an atom. As an application of the rotation
matrix, let us rotate the system (the atom) from the x, y to the x0 , y 0 axis shown in the figure. The

Figure 9.10: The Rotation of an atom described by it Y`m wave function. The dark circle marks the
same point of the atom before and after the rotation.

original system LAB has the angular wavefunction Y`m (θφ) =< r̂ | `m >, and the rotated atom
g has the angular wavefunction D Y`m (θφ) =< r̂ | D | `m > . The rotated system is described
LAB
as earlier by a new function of the old coordinates, which is physically equal to the old function of
the new coordinates:
D Y`m (θφ) = f`m (θφ) = Y`m (θ0 φ0 ),
where θ0 φ0 are the angular coordinates of the point P ( marked by the dark circle in the figure),
which is the same point of the system. Hence, using the closure property of the | `m > states and
the fact that [`2 D] = 0, we have
X
D Y` m (θφ) =< r̂ | D | ` m >= < r̂ | `m0 > < ` m0 | D | ` m >
m0

and deduce that the spherical harmonics rotate as


X
D Y`m (θφ) = Y`m (θ0 φ0 ) = `
Y`m0 (θφ) Dm 0 m (α, β, γ).

m0

in the last step we introduced the matrix representation of the operator D.


We now claim that r

`
Dm0 0 (α, β, γ) = Y ∗ 0 (β, α).
2` + 1 `m
Recall that α → φ and β → θ. To understand the reason for this assertion, note that for m = 0 our
result for the rotation becomes
r
X 4π
Y`0 (θ0 φ0 ) = Y`m0 (θφ) ∗
Y`m 0 (β, α).
2` + 1
m0
q
Since Y`0 (θ0 φ0 ) = 2`+1 0
4π P` (cos θ ), we obtain

4π X
P` (cos θ0 ) = ∗
Y`m0 (θφ) Y`m 0 (β, α).
2` + 1 0
m

This is the familiar addition theorem that relates the the θφ coordinates of a point P to its θ0 φ0
coordinates after a rotation described by the angles β, α.
Later we will use some of these properties to describe the main features of spherical tensor
operators.
This addition formula can also be obtained directly from the Wigner rotation functions (see H.W.)
CHAPTER 9. SYMMETRIES 241

9.2.14 Clebsch-Gordan Reduction


J
The matrices DM M 0 (α, β, γ) are “representations of the rotation group R(3).” to understand the
connection between these functions and the Clebsch-Gordan coefficients, let us consider the
problem of adding angular momentum and rotation frames in two equivalent ways. This will lead
us to the notion of reducing a product representation to a sum of irreducible representations.
Let us consider the addition of angular momentum and use the fact that we can vectorially add
~j1 + ~j2 in either the original x, y, z
X
ϕj1 ,m1 (ξ1 ) ϕj2 ,m2 (ξ2 ) = < J M | j1 m1 j2 m2 > ΦJM (ξ1 , ξ2 )
JM

or in the rotated x0 , y 0 , z 0
X
ϕj1 ,m01 (ξ10 ) ϕj2 ,m01 2 (ξ20 ) = < J M 0 | j1 m01 j2 m02 > ΦJM 0 (ξ10 , ξ20 )
JM 0

frames. q
We have the rotation properties
X j1
ϕj1 ,m01 (ξ10 ) = ϕj1 ,m1 (ξ1 ) Dm 1 ,m
0
1
m1

and X j2
ϕj2 ,m02 (ξ20 ) = ϕj1 ,m2 (ξ2 ) Dm 2 ,m
0 ,
2
m2

or more formally X j1
m01 > = D | j1 m01 >=
| j1g | j1 , m1 > Dm 1 ,m
0 .
1
m1

Hence X j1 j2
ϕj1 ,m01 (ξ10 ) ϕj2 ,m01 2 (ξ20 ) = ϕj1 ,m1 (ξ1 ) ϕj2 ,m2 (ξ2 ) Dm 1 ,m
0 Dm ,m0
2 1 2
m1 m2
X j1 j2
= < JM | j1 m1 j2 m2 > ΦJM (ξ1 ξ2 ) Dm 1 ,m
0 Dm ,m0
2 1 2
JM m1 m2
X
= < JM 0 | j1 m01 j2 m02 > ΦJM 0 (ξ10 ξ20 )
JM 0
X
= < JM 0 | j1 m01 j2 m02 > DM
J
M 0 ΦJM (ξ1 ξ2 )
JM 0 M

here we used X
ΦJM 0 (ξ10 ξ20 ) = J
DM M 0 ΦJM (ξ1 ξ2 )
M

in the last step. We have added angular momentum and rotated in different sequences. From the
relation X
< JM 0 | j1 m01 j2 m02 > DM
J
M 0 ΦJM (ξ1 ξ2 ) =
JM 0 M
X j1 j2
< JM | j1 m1 j2 m2 > Dm 1 ,m
0 Dm ,m0 ΦJM (ξ1 ξ2 ),
2 1 2
JM m1 m2

with the common factor of ΦJM (ξ1 ξ2 ), we arrive at


X
< JM 0 | j1 m01 j2 m02 > DM
J
M0 =
JM 0 M
X j1 j2
< JM | j1 m1 j2 m2 > Dm 1 ,m
0 Dm ,m0 .
2 1 2
JM m1 m2

q Here, ξi0 denotes the space and spin “coordinate” labels for the ith particle.
CHAPTER 9. SYMMETRIES 242

Using CG orthogonality, this yields


X j1 j2
DMJ
M0 = < JM | j1 m1 j2 m2 > Dm 1 ,m
0 Dm ,m0
2
< j1 m01 j2 m02 | JM 0 >
1 2
m1 m2 m01 m02

and
J=j
X 1 +j2 X
j1 j2 J 0 0 0
Dm 1 ,m
0 Dm ,m0 =
2
< j1 m1 j2 m2 | JM > DM M 0 < JM | j1 m1 j2 m2 > .
1 2
J=|j1 −j2 | M M 0

The last expression is the reduction formula, which reduces a product of rotation matrices to a
sum over irreducible rotations. An abbreviated notation for this reduction is
J=j
X 1 +j2
j1 j1
D ⊗D = DJ .
J=|j1 −j2 |

A familiar case is the spin 1/2 ⊗ spin 1/2 reduction to triplet J = S = 1 and singlet J = S = 0
states. Note the above reduction results all involve the same rotation R = (α, β, γ), and should not
be confused with the addition formula which involved different angles.

9.2.15 Irreducible Spherical Tensors


Let us now discuss tensor operators. It is helpful to understand the efficacy of various operators in
their ability to transfer angular momentum. Quantum mechanics involves matrix elements, i.e. it
involves initial and final states and an operator. In scattering theory, we saw that an operator has
an ability to transfer linear momentum; it also has the ability to transfer both orbital and spin
angular momentum. Matrix elements also describe transitions in atomic, nuclear and solid state
systems and we again need to know the role of operators. Other attributes such as parity and
reversal of motion can affect which matrix elements are zero, which are small and which are large.
Indeed, it is the ability to deduce selection rules from symmetries and associated conservation laws,
that is the hallmark of the “new” quantum mechanics.
So we should understand the rotation properties of operators. In the following, we start with some
simple Cartesian tensors and show how they can be reduced to classes that maintain their integrity
under rotations. That is, we form groups of tensor operator that mix among themselves under
rotations, and do not step outside of their type. We will introduce a special basis and define
spherical tensor operators (STO) which will be labelled by a rank (or an angular momentum) and
will form an irreducible spherical tensor.

Reduction of Cartesian Tensors


The process of reduction and the idea of an irreducible tensor is nicely illustrated by starting from
the transformation of a Cartesian tensor under a rotation, which is given by
X
0
Tµνλ··· = aµσ aντ aλρ · · · Tστ ρ··· ,
στ ρ···

where µνλ, στ ρ · · · , label the Cartesian coordinates x, y, z. The aµσ · · · are orthogonal 3 × 3
rotation matrices. Under rotations these tensors have a rank described by scalars (tensors of rank
0), vectors (tensors of rank 1) and tensors of rank 2 or higher.
Let us consider a Cartesian tensor of rank two Tµλ , which has nine components
Txx , Txy , Txz , Tyx , Tyy , Tyz , Tzx , Tzy , Tzz . From these nine components we can extract irreducible
tensors by the following process.
1. Form a scalar (` = 0) by taking the trace
X
Tr[T ] = Txx + Tyy + Tzz = Tµµ .
µ

This is a single number.


CHAPTER 9. SYMMETRIES 243

2. Form an anti-symmetric tensor (` = 1) as

(Tµν − Tνµ )/2,

which are three numbers.


3. Form a symmetric traceless tensor (` = 2)
1 X
(Tµν + Tνµ )/2 − δµν Tλλ ,
3
λ

which are five numbers.


Thus we reduce the nine original numbers of the Cartesian rank two tensor in three classes; namely
a scalar plus a vector, plus a symmetric, traceless tensor. These add to 1+3+5=9 components, so
we have simply just reorganized the components.
The important property of this reduction is that under rotations each of the three types transform
among themselves, i.e. the rotated symmetric trace-less tensor is a combination of the original
symmetric trace-less tensor. This property is reducibility and the final three types are called
irreducible tensors, since we can not reduce them any further. We have used the fact that under a
rotation the symmetry type does not change.
The association with angular momentum is already appearing in that the number of components is
given by the 2` + 1 rule.
Now it is more convenient to introduce a spherical basis instead of the Cartesian (x,y,z) basis to
express irreducible tensors. For example an irreducible vector or 3 -component tensor can be
expressed as r

r1µ = r Y1µ (θ, φ),
3
with
1
r1±1 = ∓ √ (x ± iy); and r10 = z.
2
~ ( such as p~, or L),
For a general vector operator A ~ the spherical components or basis are:

1
A1±1 = ∓ √ (Ax ± iAy ); and A10 = Az .
2

Law of Transformation
The irreducible spherical tensors transform as the spherical harmonics do under R = {α, β, γ}
rotations, i.e.
(λ)
X
DYλµ (θ, φ) = Yλµ (θ0 , φ0 ) = Yλµ0 (θ, φ)Dµ0 µ (α, β, γ).
µ0

The spherical tensors can be introduced as

Tλµ (r) = rλ Yλµ (θ, φ)f (| r |),

and for the transformed case

Tλµ (r0 ) = Tλµ


0
= rλ Yλµ (θ0 , φ0 )f (| r |),

and then one sees that the basic STO transformation law is
(λ)
X
0
Tλµ = Tλµ0 Dµ0 µ (α, β, γ).
µ0
q

For example, consider the rank-1 spherical tensor r1µ = 3 r Y1µ (θ, φ), we have

(λ)
X
0
r1µ = r1µ0 Dµ0 µ (α, β, γ).
µ0
CHAPTER 9. SYMMETRIES 244

Now take µ = 0, and set γ = 0 and β = θ. Using the D(1) values (see H.W.), this becomes

sin β sin β
z 0 = z cos β − e−iα √ r11 + eiα √ r1−1 = z cos β + x cos α sin β + y sin α sin β.
2 2
For α = 0, this reduces to z 0 = z cos β + x sin β, which is clearly a simple rotation by β. Thus the
spherical tensor basis transformation records the proper components, with the extra ingredient of
being irreducible.

Polarization of an STO
We should not think that all tensor operators are simply of the form f (| r |) rλ Yλ,µ (θ, φ).
Operators of other types can be produced by the scalar operator A ~ · ∇.
~ For example consider:

~ rλ Yλ,µ (θ, φ)
~σ · ∇

for λ = 1, r r
~ rY1,µ (θ, φ) = 3 ~ r1µ = 3
~σ · ∇ ~σ · ∇ σ1µ ,
4π 4π
where
1
σ1µ=±1 == ∓ √ (σx ± iσy ); and σ10 = σz .
2
Note that the STO σ1±1 is related to the raising and lowering operators σ± by σ1±1 = ∓ √1 σ± .
2
Another example of how to generate more general STOs is, for λ = 1,
r r
~ ~ 3 ~ ~ 3
L · ∇ rY1,µ (θ, φ) = L · ∇ r1µ = L1µ ,
4π 4π
where
1
L1µ=±1 = ∓ √ (Lx ± iLy ); and L10 = σz .
2
This process for generating general STO by action of the scalar operator A ~·∇~ is called
“polarization.” It can be generalized to higher rank tensor operators, which are of a “pure” type
λ
Y
Aλµ = [ ~ · ∇)
(A ~ n ] rλ Yλ,µ (θ, φ),
n=1

which is pure in that the resultant operator is formed from A only. A mixed operator, say mixed
between A and B would be generated by
NA
Y λ−N
YA
Tλµ (A, B) = [ ~ · ∇)
(A ~ nA ] [ ~ · ∇)
(B ~ nB ] rλ Yλ,µ (θ, φ),
nA =1 nB =1

which would include NA of the A operators and λ − NA of the B operators. A familiar example of
a mixed operator is orbital angular momentum as built from r1µ and p1ν , see HW.

Motivation
The physical motivation for discussing spherical tensor operators is based on the fact that
observables can be characterized as various tensor operators. For example we have:
• Scalar observables, such as the density, kinetic energy and/or the potential energy of a
system.
~ = e~r, or magnetic dipole operator
• Vector observables, such as the current, electric dipole D
µ
~ = g~σ of a system, which involves static properties and also transitions.
CHAPTER 9. SYMMETRIES 245

• Higher rank observables, such as the electric quadrupole Q2µ ∼ r2 Y2µ or magnetic
quadrupole operators or higher cases, such as octupole operators. These again can be related
to both static and transition properties of a system. These transitions can involve
electromagnetic, weak or strong interactions.
The point is we can identify many operators by their tensor rank and that tells us how they
behave with regard to transferring angular momentum consistent with conservation of the total
angular momentum of the system. The reduction of Cartesian tensors to irreducible STOs, which
was illustrated earlier, permits one to obtain operators that transform among themselves under
rotation. We were led to the following definition of a spherical tensor operator (i.e an irreducible
operator of tensor rank λ.)

(λ)
X
T̃λµ = D Tλµ D† = Tλµ0 Dµ0 µ (α, β, γ).
µ0 =−λ

The appearance of the same rank label λ on both sides of this equation is the irreducible aspect,
since only those operators are involved, albeit of different components µ, in the rotation.
An equivalent way of defining the STO Tλµ is obtained by considering an infinitesimal rotation
keeping all terms to order . We have with α0 →  → 0,
α0 ~  ~
D = e−i h̄ (~n·J) → 1 − i (~n · J),

and its matrix element version

Dµλ0 µ (α, β, γ) = δµ0 µ − i < λµ0 | ~n · J~ | λµ > .

Therefore,
 ~ Tλµ ] = Tλµ − i 
X
D Tλµ D† = Tλµ − i [~n · J, Tλµ0 < λµ0 | ~n · J~ | λµ > .
h̄ h̄ 0
µ

We conclude that another way to define an STO is by


X
~ Tλµ ] =
[~n · J, Tλµ0 < λµ0 | ~n · J~ | λµ > .
µ0

It is simpler to express the above equation using the J± operators. Picking n̂ as x̂ then ŷ, and ẑ,
one finds three commutators
X
[Jx , Tλµ ] = Tλµ0 < λµ0 | Jx | λµ > .
µ0
X
[Jy , Tλµ ] = Tλµ0 < λµ0 | Jy | λµ > .
µ0

and
[Jz , Tλµ ] = h̄µTλµ .
p
Combining the first two and using J± | λµ >= h̄ λ(λ + 1) − µ(µ ± 1) | λµ ± 1 >, we get
p
[J± , Tλµ ] = h̄ λ(λ + 1) − µ(µ ± 1) Tλµ±1 .
Summary
Spherical tensor operators are define by either

(λ)
X
D Tλµ D† = Tλµ0 Dµ0 µ (α, β, γ).
µ0 =−λ

or equivalently by p
[J± , Tλµ ] = h̄ λ(λ + 1) − µ(µ ± 1) Tλµ±1 .
and
[Jz , Tλµ ] = h̄µTλµ .
CHAPTER 9. SYMMETRIES 246

The Algebra of Spherical Tensor Operators–Racah Algebra


Tensor operators can be added, subtracted and multiplied. The text by Brink & Satchler has all
the information. Racah’s classic papers, which are in the collection “Quantum Theory of Angular
Momentum ” L.C. Biedenharn and H. Van Dam (Academic press), and Wigner’s book on group
theory are other good sources.
Addition
If we add two tensor operators Rλµ , Sλµ of the same rank and same component λµ, we get a tensor
operators Tλµ
Tλµ = Rλµ + Sλµ .
The test for a tensor operator is how it transforms. Clearly, if

(λ)
X
D Rλµ D† = Rλµ0 Dµ0 µ (α, β, γ),
µ0 =−λ

and

(λ)
X
D Sλµ D† = Sλµ0 Dµ0 µ (α, β, γ),
µ0 =−λ

then
+λ +λ
(λ) (λ)
X X
D Tλµ D† = D Rλµ D† + D Sλµ D† = (Rλµ0 + Sλµ0 )Dµ0 µ (α, β, γ) = Tλµ0 Dµ0 µ (α, β, γ).
µ0 =−λ µ0 =−λ

Thus Tλµ is an STO.


Multiplication
There are two related ways to multiply two STOs.

A. Tensor Product A tensor product is defined by using the C-G coefficient in


X
TLM = < λ1 µ1 λ2 µ2 | LM > Rλ1 µ1 Sλ2 µ2 ≡ [Rλ1 ⊗ Sλ2 ]LM ,
µ1 µ2

where the rank of the resultant operator TLM satisfies the usual C-G rule ∆(L, λ1 , λ2 ) = 0,
which means L ranges from λ1 + λ2 to | λ1 − λ2 |, is steps of one. The proof that TLM is an
STO is left as an exercise.

B. Scalar product For λ1 = λ2 = λ a scalar product can also be defined, which is related to the
above TLM with L = M = 0; namely
X (−1)λ X
[Rλ ⊗ Sλ ]00 = < λµ λ − µ | 00 > Rλµ Sλ−µ = √ (−1)µ Rλµ Sλ−µ ,
µ
2λ + 1 µ

(−1)λ
=√ (Rλ • Sλ ).
2λ + 1
(−1)λ+µ
where we used < λµ λ − µ | 00 >= √
2λ+1
. The scalar or dot product r is therefore define by


X X
(Rλ • Sλ ) ≡ (−1)µ Rλµ Sλ−µ = Rλµ Sλµ .
µ µ

We are now ready to use the transformation properties of these operators, along with the
transformation of the states to derive the important Wigner-Eckart theorem
r This ~ · S.
can be readily applied to the rank one case to recover the usual vector dot product R ~
CHAPTER 9. SYMMETRIES 247

9.2.16 The Wigner-Eckart Theorem


The Wigner-Eckart (W-E) theorem is a very important general theorem. Several procedures are
available to prove the theorem. I shall present a rather simple proof using ordinary quantum
mechanics. Other proofs, using group theory methods, are much more elegant. In addition to the
procedures described in this section, see the procedures in Messiah.
First let me state the theorem and its significance. Then it will be proven and applied to various
operators of interest.
The Wigner-Eckart theorem allows us to separate the geometrical ( or kinematical) from the
dynamical aspects of a matrix element of a tensor operator of rank λ. It states:

< α J M | Tλµ | α0 J 0 M 0 >=< J M | λµJ 0 M 0 > (α J kTλ kα0 J 0 ).

The matrix element’s dependence on orientation, i.e. on the quantum numbers associated with the
choice of ẑ axis (M M 0 , and µ) is isolated into the Clebsch-Gordan coefficient
< J M | λµJ 0 M 0 > . The factor (α J kTλ kα0 J 0 ) is independent of orientation, i.e., independent of
m−type quantum numbers; it is called a “reduced matrix element” and also a “double-barred
matrix element.” The theorem is the fact that the orientation effect is totally in the
Clebsch-Gordan coefficient. The remaining factor contains the dependence on the other quantum
numbers J, J 0 λ, and also on other quantum numbers needed to describe the state here denoted by
α and α0 . For example, α could denote the usual n` and parity quantum numbers of a state. The
“reduced matrix element” is defined in many references with an extra square root factor as in:

(α J kTλ kα0 J 0 )
< α J M | Tλµ | α0 J 0 M 0 >=< J M | λµJ 0 M 0 > √ .
2J + 1
I will use the earlier form, following the conventions of Brink & Satchler.
The theorem separates the dependence on orientation into a Clebsch-Gordan coefficient, which is
the same factor for all spherical tensor operators of the same rank λ between the same states; thus,
the angular momentum kinematics of quantum theory is isolated into the C-G coefficients. The
evaluation of the reduced matrix element is the dynamical aspect and involves a single evaluation
of the original matrix element with a convenient set of values for the M, M 0 , µ quantum numbers.
The WE theorem is partly a theorem, in the occurrence of a Clebsch-Gordan coefficient, and
partly a definition in the reduced matrix element convention.
Proof of the theorem is now presented. The W-E theorem is a consequence of the transformation
properties of the spherical tensor operator Tλ µ and of the C-G coefficient properties. First consider
the action of Tλ µ on an eigenstate of J 2 and Jz (| αJ M >) : Tλ µ | αJ M > . Here α again
denotes other quantum numbers needed to fully specify the state. We see from the commutators
defining Tλ µ that

Jz Tλ µ | αJ M >= [Jz , Tλ µ ] | αJ M > +Tλ µ Jz | αJ M >,

= h̄ (µ + M ) | αJ M > .
which shows that Tλ µ | αJ M > is an eigenstate of Jz with eigenvalue h̄ (µ + M ). Also consider:

J± Tλ µ | αJ M >= [J± , Tλ µ ] | αJ M > +Tλ µ J± | αJ M >,


p p
= λ(λ + 1) − µ(µ ± 1) | αJ M > + J(J + 1) − M (M ± 1) | αJ M ± 1 > .
This equation suggests that one can think of Tλ µ | αJ M > as a product state of the type
| λ µ > | αJ M >, involving an addition of angular momentum as if ~λ + J. ~ This formal
~
identification suggests that we consider the C-G combination (adding λ to J~ )
X
| ξJ 0 M 0 ] ≡ < J 0 M 0 | λ µ J M > Tλ µ | αJ M >,
µM

where ξ denotes the quantum number λ, J and α. The above linear combination is given a name
| ξJ 0 M 0 ] with the square bracket | · · ·] used to indicate that a proof is needed to show that this
CHAPTER 9. SYMMETRIES 248

particular combination is indeed an eigenstate of J 2 and Jz . We shall replace | ξJ 0 M 0 ] →| ξJ 0 M 0 >


only after it is shown that
J 2 | ξJ 0 M 0 ] = h̄2 J 0 (J 0 + 1) | ξJ 0 M 0 ],
and
Jz | ξJ 0 M 0 ] = h̄M 0 | ξJ 0 M 0 ].
Obviously, we have
X
Jz | ξJ 0 M 0 ] = < J 0 M 0 | λ µ J M > Jz Tλ µ | αJ M >
µM
X
= < J 0 M 0 | λ µ J M > h̄(µ + M ) Tλ µ | αJ M >= h̄M 0 | ξJ 0 M 0 ].
µM

So we just need to demonstrate the J 2 eigenvalue result.


The fact that | ξJ 0 M 0 ] is an eigenstate of J 2 with eigenvalue J 0 (J 0 + 1)h̄2 can be demonstrated in
various ways. One way is to use J 2 = Jz2 + 21 (J+ J− + J− J+ ), and to show that it suffices to have
p
J± | ξJ 0 M 0 ] = h̄2 J 0 (J 0 + 1) − M 0 (M 0 ± 1) | ξJ 0 M 0 ± 1],
to establish that this state is an eigenstate of J 2 . This is left to the reader.
Another way is to show that | ξJ 0 M 0 ] has the correct rotation property, which follows from
X
D | ξJ 0 M 0 ] = < J 0 M 0 | λ µ J M > DTλ µ D† D | αJ M >,
µM

(λ) (J)
X X
= < J 0 M 0 | λ µ J M > Dµ00 µ DM 00 M Tλ µ00 | αJ M 00 >,
µM µ00 M 00
(λ) (J)
X X X
= {< J 0 M 0 | λ µ J M > Dµ00 µ DM 00 M < λ µ00 J M 00 | J M >}Tλ µ00 | αJ M > .
J M µM µ00 M 00

Now the combination in the {· · ·} brackets is in the form of the reduction formula
(λ) (J) (J 0 )
X X
{< J 0 M 0 | λ µ J M > Dµ00 µ DM 00 M < λ µ00 J M 00 | J M >} = δJ J 0 DMM ,
µM µ00 M 00

hence
(J 0 )
X
D | ξJ 0 M 0 ] = DMM | ξJ 0 M]
M
0 0
The last equation demonstrates that | ξJ M ] rotates as an eigenstate of J 2 with eigenvalue
J 0 (J 0 + 1)h̄2 . Therefore, we can now apply the state designation and make the replacement
| ξJ 0 M 0 ] →| ξJ 0 M 0 > .
Having shown that | ξJ 0 M 0 > to be an eigenstate, we now proceed to the final steps of the proof of
the W-E theorem. Consider:
X
| ξJ 0 M 0 >≡ < J 0 M 0 | λ µ J M > Tλ µ | αJ M >,
µM

and recall that ξ denotes the λ, α, J quantum numbers. Using closure we have
X X
| ξJ 0 M 0 >= | α00 J 00 , M 00 >< α00 J 00 , M 00 | ξJ 0 M 0 >= | α00 J 0 , M 0 >< α00 J 0 , M 0 | ξJ 0 M 0 >,
α00 J 00 ,M 00 α00

where we used the eigenstate property of | ξJ 0 M 0 > and < α00 J 00 , M 00 |; namely that
< α00 J 0 , M 0 | ξJ 0 M 0 >∼ δJ 0 J 00 δM 0 M 00 . The overlap < α00 J 0 , M 0 | ξJ 0 M 0 > can be shown to be
independent of M 0 , s and we can designate is to be a function
< α00 J 0 , M 0 || ξJ 0 M 0 >→ (α00 J 0 kTλ k αJ),
s 00 0 0 0 0
p of J+ on a bra and a ket: < α J , M | J+ | ξJ M − 1 >=
p This independence follows from considering the action
J 0 (J 0 + 1) − (M 0 − 1)M 0 < α00 J 0 , M 0 | ξJ 0 M 0 >= J 0 (J 0 + 1) − M 0 (M 0 − 1) < α00 J 0 , M 0 − 1 | ξJ 0 M 0 − 1 > .
Hence < α00 J 0 , M 0 | ξJ 0 M 0 >=< α00 J 0 , M 0 − 1 | ξJ 0 M 0 − 1 > which demonstrates that < α00 J 0 , M 0 | J+ ξJ 0 M 0 > is
independent of M 0 .
CHAPTER 9. SYMMETRIES 249

where we have included the ξ → λ, α, J relationship.


Now from X
| ξJ 0 M 0 >= < J 0 M 0 | λ µ J M > Tλ µ | αJ M >,
µM

we form
X
< α00 J 0 M 0 | ξJ 0 M 0 >= < J 0 M 0 | λ µ J M > < α00 J 0 M 0 | Tλ µ | αJ M >,
µM

hence X
(α00 J 0 kTλ k αJ) = < J 0 M 0 | λ µ J M > < α00 J 0 M 0 | Tλ µ | αJ M > .
µM

From the orthogonality of the C-G above, we arrive at the final result

< α0 J 0 M 0 | Tλµ | αJM >= (J 0 M 0 | λµJM )(α0 J 0 k Tλ k αJ).

Outline of the Group Theory Proof of the Wigner-Eckart Theorem


The group theory approach to the W-E theorem starts from the properties of states and STOs
under rotations:
(J)
X
D | JM >= | JM 0 > DM 0 M ,
M0
(λ)
X
D Tλµ D† = Tλµ0 Dµ0 µ .
µ0
(J) (J)∗
Note that: DM M 0 =< JM | D | JM 0 >, and DM M 0 =< JM 0 | D† | JM > .
Using these relations we have:

< α0 J 0 M 0 | Tλ µ | αJM >=< α0 J 0 M 0 | D† (DTλ µ D† ) D | αJM >,


from the unitary property of D† D = 1 = DD† . Hence
X
< α0 J 0 M 0 | Tλ µ | αJM >= < α0 J 0 M 00 | Tλµ0 | αJM 000 >
M 00 M 000 µ0

(J 0 )∗ (λ) (J)
×DM 00 M 0 (αβγ) Dµ0 µ (αβγ) DM 000 M (αβγ).
This holds true for all rotations R = (αβγ). t
We can therefore integrate over these Euler angles using the group integration
Z 2π Z π Z 2π
dα sin βdβ dγ,
0 0 0

on both sides. On the right this yields


Z
dα sin βdβ dγ < α0 J 0 M 0 | Tλ µ | αJM >= (2π)2 2 < α0 J 0 M 0 | Tλ µ | αJM >,

and on the left X


< α0 J 0 M 00 | Tλµ0 | αJM 000 >
M 00 M 000 µ0
Z
(J 0 )∗ (λ) (J)
× dα sin βdβ dγDM 00 M 0 (αβγ) Dµ0 µ (αβγ) DM 000 M (αβγ).

(J 0 )∗ 00 0 (J 0 )
Note that DM 00 M 0 (αβγ) = (−1)M −M D−M 00 −M 0 (αβγ).
An important result of the rotation group is now useful
Z   
(C) A B 2 A B C A B C
dα sin βdβ dγDcc0 (αβγ) Daa0 (αβγ) Dbb0 (αβγ) = 8π
a b c a0 b0 c0
t We are using α in two distinct contexts; one is as an Euler angle within the Wigner D functions, the other as a

set of state quantum numbers n`. These should not be confused.


CHAPTER 9. SYMMETRIES 250

see B&S page 148. The 3-j’s are related to the C-G’s by

 
j1 −j2 +M j1 j2 J
< j1 m1 j2 m2 | J − M >= (−1) 2J + 1
m1 m2 M

Combining all of the above, one gets that

< α0 J 0 M 0 | Tλ µ | αJM >=< λ µ J M | J 0 M 0 > ×


0
X
0 0 00 000 0 000 0 (−1)2(J−M −λ)
00
<αJ M |T λµ0 | αJM >< λ µ J M |J M > .
2J 0 + 1
M 00 M 000 µ0

The last factor is an average over all m− type projections and is therefore independent of
orientation and can be replaced by the reduced matrix element notation. Hence we again arrive at:

< α0 J 0 M 0 | Tλ µ | αJM >=< λ µ J M | J 0 M 0 >< α0 J 0 k Tλ kαJ > .


Chapter 10

Quantum Dynamics

10.1 Stationary State Perturbation Theory


Reference: Messiah Chapter XVI.
The stationary states are determined by solving the time independent Schrödinger equation

H | ψn >= En | ψn >,

where a stationary state satisfies


i
| Ψ(t) >= e− h̄ En t | ψn > .

The eigenvalues form a discrete plus a continuum spectrum along with the corresponding set of
eigenfunctions. The spectrum of eigenvalues can be degenerate for certain energies; by degenerate
we mean that two or more distinct solutions, | ψ1n >, | ψ2n > · · · , exist having the same energy
eigenvalue. For a central force problem, one has a 2` + 1 fold degeneracy due to energy being
independent of orientation. Another example of energy degeneracy occurs in the 2s and 2p levels
for a Coulomb problem.
Aside from a few cases, such as the harmonic oscillator and the Coulomb interaction, one can not
solve for the exact eigenfunctions and eigenvalues. It is necessary to use approximation methods,
which we shall now discuss.
First it is assumed that the full Hamiltonian can be separated into two parts

Hg = H0 + gV,

where the parameter g is allowed to vary, with limg→1 Hg = H, the full Hamiltonian. The
“coupling constant” is introduced to give us control over the perturbing term (gV ). Also, the
limits limg→1 Hg = H, and limg→0 Hg = H0 , are assumed to exist and we also assume usually that
one can turn the interaction on and off smoothly, with the corresponding eigenvalues and
eigenfunctions also behaving in a smooth or “analytic” manner with respect to changing g.
Typically, H0 is the kinetic energy operator and V is the potential energy V term which falls of
with distance as 1/r or faster. Another common choice is to take H0 to be the Hamiltonian for an
p2 p2 2
harmonic oscillator H0 = 2m + 21 mω 2 r2 , or a Coulomb problem H0 = 2m − Zer , which are then
perturbed by an interaction V. Sometimes an explicit separation can not be conceived and one
invokes the idea of a model Hamiltonian by taking

H = Hmodel + V − Hmodel = Hmodel + “V ”,

where it is hoped that Hmodel simulates the main dynamics and that the residual interaction “V ”
is a perturbation. This approach is the basis for much of atomic and nuclear and also solid state
physics calculations.
Let us introduce suitable notation. To designate the eigenstates of H we write

| αEn >=| ψn > with H | ψn >= En | ψn >,


CHAPTER 10. QUANTUM DYNAMICS 252

where α refers to the eigenvalues of the additional operators, which along with H form a complete,
commuting, compatible set of operators. ( We shall later also need to label the various states as
H | ψin >= Ein | ψin >, where i is a label for the degeneracy i = 1 · · · Dn . For clarity, we will
suppress this degeneracy label and consider nondegenerate (i = 1) cases first. The degenerate case
will be discussed later.)
States for the Hamiltonian H0 are designated by

| βεin >=| ϕin > with H0 | ϕin >= εin | ϕin >,

where β refers to the eigenvalues of the additional operators which along with H0 form a complete,
commuting, compatible set of operators, such as `m` sms . The label i is used again to denote the
various degenerate states with the same energy ε1n , ε2n · · · εdn n . Both discrete (bounded) and
continuum state eigenvalues are included. The eigenstates of H0 are a complete orthonormal set
which provide basis states for treating the effect of the interaction V .

10.1.1 The Secular Determinant


The set of states | ϕn >,a can be used to express the stationary state Schrödinger equation in
matrix form:
< ϕn | H0 + gV | ψ >= E | ψ >,
X
< ϕn | H0 − E + gV | ϕn0 > < ϕn0 | ψ >= 0.
n0

We define the matrix Ω by

Ωnn0 (E) ≡< ϕn | H0 − E + gV | ϕn0 >= δnn0 (εn − E) + g < ϕn | V | ϕn0 >,

and hence X
Ωnn0 (E) < ϕn0 | ψ >= 0, .
n0

The above equation is of the matrix form Ax = 0, and hence has a solution only if the determinant
of |ω is zero. Therefore, the various eigenvalues En are obtained from the secular determinant
equation det Ω(E) = 0, or

ε1 − E + gV11 gV12 gV13 · · ·

gV21 ε2 − E + gV22 gV23 · · ·
· · = 0


gV 31 gV32 ε 3 − E + gV33 ·
. .
··· .

The sum over n in general includes both discrete and continuum states. In view of that
complication, and also because of the usually large number of discrete states, solution of the above
secular equation is often totally impracticable. It is necessary to reduce the size of the problem;
here we have an N × N matrix Ω with N very large, indeed infinite. One important way to reduce
the problem’s size is to make use of the symmetry properties of H.
Note that after solving det Ω(E) = 0 for various En eigenvalues, one can also determine the wave
function.

10.1.2 Symmetry and Reduction of the Secular Equation


If H has a symmetry, then the secular equation can be reduced considerably. It is the reduction of
the secular “matrix” to smaller sub-matrices that, to a large extent, motivates our search for
symmetries. Group theory provides a means for formulating general symmetry properties and
serves hence as a means of reducing the size of the problem.
Let me illustrate how a symmetry can be used to reduce the problem of solving the secular
equation det Ω(E) = 0. Recall that our states include other quantum numbers labelled by β, i.e.
we have states | ϕn >=| βεn > . The quantum numbers β could for example be the total angular
a The degeneracy index i is suppressed.
CHAPTER 10. QUANTUM DYNAMICS 253

momentum h̄2 J(J + 1) and the associated Jz quantum number-β : JM. In addition, β could
denote the parity of the state or some other “good” quantum numbers. b
For simplicity, let us assume that β is simply the JM quantum numbers. If it is known that V
preserves this symmetry property of H0 , it follows that the secular matrix is reduced to
sub-matrices. From [J, ~ V ] = 0, it follows that states with different values of β can not be coupled
by V and hence Ω(E) becomes split or block-diagonal; for example

Ωβn,β 0 n0 = δJJ 0 δM M 0 [(εn − EJ )δnn0 + < nJM | V | n0 JM.] = δJJ 0 δM M 0 ΩJnn0 (EJ ).

The big Ω matrix reduces to sub-matrices ΩJ for each value of J:


J
Ω 1 0 0 0
J2
0 Ω 0 0
0 = 0

0
0 Ω J3
. .
0 0 0 .

Since the number of states with angular momentum J might be small, the solution of the separate
secular equations det ΩJ (EJ ) = 0 may become manageable. Hence, the rotational symmetry has
been used to partly diagonalize Ω and reduce the size of the secular equations.

Perturbation Theory
Further reduction of the secular equation can be made using addition symmetries such as parity,
isospin , etc. Nevertheless, it is still usually too difficult to solve the secular equation because many
states and hence large matrices are involved. Another simplification occurs however if gV can be
treated as “small;” namely, one can expand the secular determinant in various ways. “Small” will
be defined later.
The approximate evaluation of det Ω(EJ ) = 0 can be formulated in various ways. Historically,
perturbation theory was developed to treat the gravitation perturbations of one planet on the
motion of others–hence the phraseology “secular determinant.” Even in that classical problem it
was necessary to approximate the evaluation of the secular equation. The techniques developed
then were simple adopted for the corresponding quantum dynamics problem.
For the simplest cases, we can directly obtain an estimate of the effect of a small gV on the
eigenvalues. Note that from the secular determinant expression above, it follows that for g = 0,
det Ω = (ε1 − E) (ε2 − E) · · · , and hence En = εn , as expected. If we neglect the off-diagonal terms
V12 , V13 · · · V21 , V23 , · · · (thereby dropping all terms of order g 2 and higher), then

det Ω = (ε1 − E + gV11 ) (ε2 − E + gV22 ) · · · = 0.

The first order perturbation theory result is then

En − εn + g < n | V | n > First Order.

Here, we denote | ϕn > simply by | n > .


The problem now is to determine the effect of the neglected “off-diagonal” terms. To include these
terms in evaluating det Ω = 0, a systematic approach is needed as presented next.

b Good quantum numbers correspond to additional eigenvalue problems and hence attributes that have zero uncer-

tainty. They are simultaneously diagonal with the Hamiltonian and form a complete set of commuting, compatible
observables.
CHAPTER 10. QUANTUM DYNAMICS 254

Non-degenerate Brillouin-Wigner Perturbation Series


Let us return to the original form of the Schrödinger equation

(E − H0 ) | ψ >= gV | ψ > .

Using < ϕn | on this equation, one finds

((E − εn ) < ϕn | ψ >= g < ϕn | V | ψ > .

and hence
< ϕn | V | ψn >
E n = εn + g ,
< ϕn | ψn >
which is an exact result since the exact wave function | ψn > appears on the right hand side (RHS).
Now we assume that each state | ϕn > is nondegenerate; for each energy εn only one state occurs.
Also, the energy En and the wave function | ψn > are assumed to evolve smoothly from each | ϕn >
and εn , hence the appropriate label “n” is introduced on the exact eigenvalue and eigenstate.
We shall assume a bound state spectrum only and thus the states can be normalized. It is
convenient to normalize our state using < ϕn | ψn >= 1, which means that the state | ψn > is not
properly normalized < ψn | ψn >6= 1; nevertheless we can renormalize | ψn > later.
With this normalization of < ϕn | ψn >= 1, we have the exact relation

En = εn + g < ϕn | V | ψn > .
With | ψn >∼| ϕn >, the first order result is recovered. A better estimate for | ψn > will provide
the means to include the aforementioned off-diagonal terms.
To get a better estimate for | ψn >, we can write | ψn > in a form analogous to that in scattering
theory
X | ϕn0 > < ϕn0 | gV | ψn >
| ψn >= 1 | ϕn > +
0
E n − ε n0
n 6=n

Note the normalization < ϕn | ψn >= 1 has been incorporated since < ϕn | ϕn0 >= δnn0 . Also
note that the exact wavefunction and eigenvalue appears on the RHS; thus, this is really an
integral equation form rather than a solution. The above expression can be shown to satisfy the
the Schrödinger equation
X
(En − H0 ) | ψn >= (En − H0 ) | ϕn > + | ϕn0 > < ϕn0 | gV | ψn >=
n0 6=n

(En − H0 − < ϕn | gV | ψn >) | ϕn > +1gV ψn >= gV | ψ >,


P P
using closure 1 = n0 | ϕn0 > < ϕn0 | = n0 6=n | ϕn0 > < ϕn0 | + | ϕn > < ϕn | .
In terms of the following projection operators
X
Pn ≡| ϕn > < ϕn | and Qn ≡ 1 − Pn = | ϕn0 > < ϕn0 | ,
n0 6=n

we write
Qn
| ψn >= 1 | ϕn > + gV | ψn > .
En − H0
A series can be obtained by iteration; one has
Qn Qn Qn
| ψn >= 1 | ϕn > + gV | ϕn > + gV gV | ϕn >
En − H0 En − H0 E n − H0
Qn Qn Qn
+ gV gV gV | ϕn > · · ·
E n − H0 En − H0 En − H0
A compact version of this Brillouin-Wigner series for the wave function is then obtained:
∞  p
X Qn
| ψn >= gV | ϕn >,
p=0
En − H0
CHAPTER 10. QUANTUM DYNAMICS 255

and for the energy


∞  p
X Qn
En = εn + g p+1 < ϕn | V V | ϕn >,
p=0
E n − H0

These two equations for exact En and exact | ψn > are the Brillouin-Wigner series. Note that the
exact energy and exact eigenfunction occurs on both sides of the equation so an additional step is
required before one can evaluate En and | ψn > . ( One possible procedure is to adjust En so that
the linear LHS crosses the RHS function evaluated to some order p.)

10.1.3 Rayleigh-Schrödinger Perturbation Theory


The usual procedure is to introduce approximate eigenvalues into the right hand side of the B-W
series and then expand the energy denominators. Doing this one obtains the Rayleigh-Schrödinger
(R-S)series. c
The R-S terms to order g 2 for | ψn > are obtained by taking En ' εn + g < ϕn | V | ϕn >, in the
energy denominators and expanding keeping terms to O(g 2 )
X | ϕn0 >< ϕn0 | gV | ϕn >
| ψn >' 1 | ϕn > +
εn + g < ϕn | V | ϕn > −εn0
n0 6=n

X X | ϕn0 >< ϕn0 | gV | ϕn00 > < ϕn00 | gV | ϕn >


+ + ···
εn − εn0 εn − εn00
n0 6=n n 6=n
00

or

X | ϕn0 >< ϕn0 | V | ϕn > X | ϕn0 >< ϕn0 | V | ϕn >


| ψn >' 1 | ϕn > +g −g 2 (< ϕn | V | ϕn >)
0
εn − εn0 0
(εn − εn0 )2
n 6=n n 6=n

X X | ϕn0 >< ϕn0 | gV | ϕn00 > < ϕn00 | gV | ϕn >


+g 2 + O(g 3 )
ε n − ε n 0 ε n − ε n00
n0 6=n n 6=n
00

This expression for | ψn > yields a series for En , in which terms to order g 3 are kept

X |< ϕn0 | V | ϕn >|2 X |< ϕn0 | V | ϕn >|2


En = εn +g < ϕn | V | ϕn > +g 2 − g3 (< ϕn | V | ϕn >)
0
ε n − ε n0 0
(εn − εn0 )2
n 6=n n 6=n

X X < ϕn | V | ϕn0 > < ϕn0 | V | ϕn00 > < ϕn00 | V | ϕn >
+g 3 + O(g 4 )
(εn − εn0 )(εn − εn00 )
n0 6=n n 6=n
00

A briefer version is made possible by introduction of the projection operator Qn and the “energy
denominator” or free propagator
1 1 1 δn0 n00
= ; < ϕn0 | | ϕn00 >= .
en ε n − H0 en ε n − ε n0
The compacted R-S series is then
Qn Qn
| ψn >'| ϕn > +g V | ϕn > −g 2 2 V | ϕn > (< ϕn | V | ϕn >)
en en
Qn Qn
+g 2 V V | ϕn > +O(g 3 )
en en
and
c The R-S series originates in the wave mechanics formulated by Schrödinger (Ann.d. Physik 80 437 (1926) ibid

79 361 (1926)and 81 109 (1926)) who based his approach on the perturbation methods by Lord Rayleigh for acoustic
and general wave problems
CHAPTER 10. QUANTUM DYNAMICS 256

Qn
En ' εn + g < ϕn | V | ϕn > +g 2 < ϕn | V V | ϕn >
en
Qn Qn Qn
+g 3 < ϕn | V V V | ϕn > −g 3 < ϕn | V 2 V | ϕn > (< ϕn | V | ϕn >) + O(g 4 )
en en en
The term −g 3 < ϕn | V Q e2n V | ϕn > (< ϕn | V | ϕn >) is called an “unlinked term,” with the other
n

terms all “linked.” This refers to later usage of diagrams to display perturbation theory. The role
of unlinked diagrams for higher orders will be discussed later.

Some Comments on R-S series


The above result has great applicability. Two features are worth noting before we turn to
applications and extensions. First, one obtains the energy to order g p+1 if the wave function is
known to lower order g p . Thus, good results for energies, and hence spectra , can be obtained with
only fair wave functions. Clearly, we need to have a reasonable wave function only in the range of
the potential V (r) to get a good estimate for the energies. Other observables, such as magnetic
moments and electromagnetic transitions, are more sensitive to the wave function and therefore
require better estimates for the wave function than is needed for evaluation of En .
Another general remark about this R-S series follows by considering the ground state energy ε0 ;
then
X |< ϕ0 | V | ϕn0 |2 >
E0 = ε0 + g < ϕ0 | V | ϕ0 > +g 2 .
0
ε 0 − ε n0
n >0

The second order term lowers the energy of the ground state because the energy denominator
ε0 − εn0 is negative.
The convergence of the R-S series depends not only on the coupling constant g, but also on the
ratio < n | V | n0 > /(εn − εn0 ). Roughly speaking, to get convergence one needs to have for n0 6= n

g < n | V | n0 >
| | 1,
(εn − εn0 )

which define “small” in the expansion. This smallness can be accomplished by a large energy
difference (εn − εn0 ) and small off-diagonal matrix elements < n | V | n0 > . Note that a small
energy difference (εn − εn0 ) can yield a non-convergent series. Indeed, for a degeneracy that we
have not accounted for, the limit (εn − εn0 ) → 0 appears and the series blows up. We need to avoid
that and consider next the case of a degenerate spectrum.
The R-S result presented here is valid only for non-degenerate and bounded levels. We can now
generalize this result to include the case of degenerate levels.

10.2 Degenerate Bound State Perturbation Theory


Degenerate levels often occur and the previous treatment must be revised. Again, degenerate levels
correspond to having several solutions for the unperturbed Hamiltonian H0 | ϕin >= εin | ϕin >,
all of the solutions | ϕ1n >, | ϕ2n >, · · · | ϕdn n > having a common energy
εn ≡ ε1n = ε2n = · · · = εdn n . The reason for our concern about the previous rendition is that we
can obtain matrix elements of the type

< 1n | V | 2n >
→ ∞,
(εn − εn )

which is hardly small, unless we prevent such terms from occurring in our expansion. The
prevention of the zero energy denominators is our present task. In fact, we must be careful even if
the levels are almost equal ( called near degeneracy) since then the associated matrix elements

< n | V | n0 >
→ Large
(εn − εn0 )
CHAPTER 10. QUANTUM DYNAMICS 257

occur and one can not use the previous expansion. d


The basic idea needed to handle degenerate or near-degenerate levels is to first remove the
degeneracy by treating part of the interaction and then using the previous development to treat
the remaining part of the interaction. This remark will be made clearer later as we proceed to: (1)
remove the degeneracy by first solving the secular equation for the dn × dn degenerate space of the
solutions | ϕin >; (2) Include other states | ϕn0 6=n > using the Rayleigh-Schrödinger perturbation
method. Here we assume for convenience that only the levels labelled by n have a dn − fold
degeneracy. The other levels n0 6= n are assumed to be non-degenerate.
The above steps are carried out by introducing projection operators which are generalized versions
of our earlier case:
dn
X X
Pn = | ϕin >< ϕin |; Qn = 1 − Pn = | ϕn0 >< ϕn0 |,
i=1 n0 6=n

assuming only the nth level is degenerate. The closure property here is
dn
X X
Pn + Qn = 1 = | ϕin >< ϕin | + | ϕn0 >< ϕn0 | .
i=1 n0 6=n

Note that Pn Pn = Pn , and Qn Qn = Qn , and Pn Qn = Qn Pn = 0, since the | ϕin >, | ϕn0 > form a
complete orthonormal set of bound state wave functions. They are eigenstates of the Hamiltonian
H0 . We seek the exact solutions of

(H0 + gV ) | ψjn >= Ejn | ψjn >

where j = 1 · · · dn counts the exact state | ψjn > and for the other n0 6= n levels

(H0 + gV ) | ψn0 >= En0 | ψn0 > .

εn00 ϕn00 hhhh


hhh En00 ψn00
h

E n0 ψ n0

εn0 ϕn0 E4n ψ4n

E3n ψ3n

 E2n ψ2n
χ4n ϕ4n 
(((
( ((
(
(
χ1n ϕ1n ` ```
```` E1n ψ1n
- g→1

Figure 10.1: This figure shows how the exact solutions typically evolve from the original solutions
for H0 as the perturbation is turned on. It is assumed that only a set of dn = 4 levels are degenerate
for a particular n with all other levels n0 , n00 nondegenerate.

As the interaction is turned off g → 0, one gets

lim Ejn = εn lim En0 6=n = εn0 6=n .


g→0 g→0

d A near degeneracy in a system can also afford remarkable advantages. For example, the mass degeneracy of the

K0 and K̄0 mesons amplifies the difference between them and serves to reveal their violation of time reversal, which
ultimately accounts for the dominance of matter over anti-matter in the universe, as explained originally by Cronin,
Fitch and Sakarov.
CHAPTER 10. QUANTUM DYNAMICS 258

For the wave functions, the usual limit occurs for the nondegenerate n0 6= n levels
limg→0 | ψn0 >=| ϕn0 >, but degenerate level n reduces to some linear combination of the original
degenerate basis, where the limit for the degenerate levels is typically of the form
dn
X
lim | ψjn >= cij | ϕin >≡| χ0jn >,
g→0
i=1

where j = 1 · · · dn labels the dn distinct linear combinations of the | ϕin >’s, which are formed to
remove the degeneracy. The process is illustrated in the figure.
Having introduced this notation, we can proceed to derive expressions for the energy and wave
functions Ejn and | ψjn > . The steps are to introduce the P and Q projection operators

[H0 + g(Pn + Qn ) V (Pn + Qn )] | ψjn >= Ejn | ψjn >,

which can be rewritten as

[Ejn − H0 − gPn V Pn ] | ψjn >= g[Pn V Qn + Qn V Pn + Qn V Qn ] | ψjn >,

The LHS above is the effect of the potential in the degenerate dn × dn space, while the RHS
includes the potential effects in the rest of the space. The LHS suggests that we first solve the
homogeneous equation
[Ẽjn − H0 − gPn V Pn ] | χjn >= 0
for the eigenvalues Ẽjn and the eigenfunctions | χjn > . This part of the process consists of solving
for the effect of the potential within the degenerate level dn × dn space. The above equation can be
solved in various ways; for example, an exact solution can often be used if the dn × dn space is not
too large. the secular determinant method is then used with

< ϕin | [Ẽjn − H0 − gPn V Pn ] | χjn >= 0


or
dn
X dn
X
[(Ẽjn − εn ) δii0 − g < ϕin | V | ϕi0 n > ] < ϕi0 n | χjn >= Ωii0 (Ẽjn ) < ϕi0 n | χjn >= 0
i0 =1 i0 =1

where we used Pn | ϕin >=| ϕin > . Therefore, solution of the dn × dn secular equation
det Ω(Ejn ) = 0, yields the eigenvalues Ẽjn and the wave functions | χjn > . (Here
< ϕi0 n | χjn >→< ϕi0 n | χ0jn >= cji ; see earlier descriptive comments.) It is assumed that the
eigenvalues Ẽjn j = 1, · · · dn are non-degenerate; i.e. that Pn V Pn , the interaction among the dn
degenerate levels has removed or “lifted” the degeneracy.
Ẽ4n χ4n

Ẽ χ
 3n 3n
( (Ẽ2n χ2n
((
χ04n ϕ4n (
(
(
(
χ01n ϕ1n
```
``` Ẽ
`` 1n χ1n
- Pn V Pn

Figure 10.2: The degeneracy is lifted by turning on the Pn V Pn interaction,


CHAPTER 10. QUANTUM DYNAMICS 259

If some of the Ẽjn s are still degenerate, then more steps would be required to obtain an initial
situation with non-degenerate levels. The judgment depends on the circumstances in a given
problem.
Sometimes the degeneracy can be removed using the first order solution of the secular equation in
the dn × dn space; then the non-degenerate values Ẽjn are obtained from
Ẽjn = εn + g < εjn | V | εjn > .
Assuming that the interaction Pn V Pn in the dn × dn space has been treated by the above steps
and that the eigenvalues Ẽjn are non-degenerate, with solutions | χjn > also known , then the rest
of the interaction can be considered. We have

< χjn | [Ejn − H0 − gPn V Pn ] | ψjn >= g < χjn | [Pn V Qn + Qn V Pn + Qn V Qn ] | ψjn >,

or using < χjn | Qn = 0, since < χjn | ϕn0 >= 0 for n0 6= n, we have

< χjn | [Ejn − Ẽjn ] | ψjn >= g < χjn | V Qn | ψjn >,

and
g < χjn | V Qn | ψjn >
Ejn = Ẽjn + .
< χjn | ψjn >
The first term Ẽjn is the result of including the interactions among the degenerate levels Pn V Pn .
The second term includes the interaction of the degenerate levels with the non-degenerate ones.
The above result is a generalization of our earlier expression En = εn + g < ϕn | V | ψn > . Now we
have treated the interaction among the non-degenerate levels first and hence obtain the
non-degenerate levels Ẽjn due to Pn V Pn only. The effect of the other levels is obtained by finding
better representation for Qn | ψjn > .
Note that iteration of the type | ψjn >=| χjn > + · · · , leads us back to

Ejn = Ẽjn + g < χjn | V Qn | χjn >= Ẽjn

since Qn | χjn >= 0. We have invoked the normalization < χjn | ψjn >= 1. We need to obtain
Qn | ψjn > to improve our result for Ẽjn ; we take (using Qn Pn = 0 )

Qn [Ejn − H0 − gPn V Pn ] | ψjn >= g Qn [Pn V Qn + Qn V Pn + Qn V Qn ] | ψjn >

or
[ Qn V Pn + Qn V Qn ] Qn V
Qn | ψjn >= g | ψjn > = g | ψjn > .
Ejn − H0 Ejn − H0
Therefore, iteration yields
Qn
Ejn = Ẽjn + g 2 < χjn | V V | χjn > + · · ·
Ejn − H0
X |< χjn | V | ϕn0 >|2
= Ẽjn + g 2 + ···
n0 6=n
Ẽjn − εn0

Now there is no zero denominator problem since Ẽjn 6= εn . ( If this turns out not to be true, that
is bad luck and you need to remove this induced degeneracy and hope that your luck improves.)
CHAPTER 10. QUANTUM DYNAMICS 260

10.3 Application of Time-independent Perturbation Theory


10.3.1 The Fine Structure of Hydrogen
References: Messiah Chapter XVI §II p 698. see also Gottfried p365, Sakurai p 304 ff. Also the
classic sources are: Condon & Shortley, Pauling & Wilson and Bethe & Salpeter Handbuch der
Physik
Our discussion of bound state perturbation theory can now be applied to the basic problem of the
Hydrogen atom. It is here that we see the ability of quantum mechanics to go beyond the old Bohr
model and the efforts of Sommerfeld to get the detailed spectrum of Hydrogen. Later, transition
rules for electromagnetic transitions between the levels of Hydrogen will be considered and
predictions for the intensity of spectral lines will be obtained. The study of Hydrogen leads
naturally to many-electron atoms and to many-body methods.
One important aspect of the corrections to the Hydrogen atom that we now consider, is that the
Dirac equation includes much of this Physics in a concise and elegant matter. Indeed, the results of
this section can be derived from the Dirac equation. The remaining discrepancy with the spectrum
of Hydrogen led the a further development, also mainly by Dirac, of relativistic field theory or
Quantum Electrodynamics (QED).
The non-relativistic unperturbed Hamiltonian for Hydrogen is simple

p2 e2
H0 − − , (10.1)
2µ r

for which we have exact eigenvalues and wave functions ( see material from last term )

H0 | ϕn > = εn | ϕn > (10.2)


µc2 1 ε0
εn = −α2 = 2
2 n2 n
ε0 = −13.6 e.v. n = 1, 2, 3, · · ·

where µ is the reduced electron-proton mass which is approximately the electron mass and
e2
α = h̄c = (137.04)−1 is the Sommerfeld “ fine structure” constant. (Sommerfeld in his introduction
of elliptical orbits and relativistic corrections originally introduced this quantity α.)
The solution for the wave function are determined by solving the radial wave equation

h̄2 d2 e2 h̄2 `(` + 1)


un` (r) + [ε n + − ]un` (r) = 0, (10.3)
2µ dr2 r 2µ r2

where < ~r | ϕn >= Rn` (r) Y`m` (r̂) = un`r(r) Y`m` (r̂). When we include the spin-states of the
electron, the wave function can be written as

ϕn`m` sms (~r) = Rn` (r)Y`m` | sms >=< ~r | ϕn`m` sms > . (10.4)

(Note: existence of a magnetic moment of the proton and the proton’s finite size is ignored
here-see Homework on the proton’s finite size effect.)
Solution of the radial equation can be carried out using the procedures described last term. That
is one subjects un` (r) to the boundary conditions un` (r = 0) = 0 and un` (r = ∞) = 0, and after
expansion in power series the series needs to be truncated by taking only integer values of
n = 1, 2, · · · . The resulting spectrum is labelled by the usual convention of s, p, d, f · · · for
` = 0, 1, 2, 3, · · · and with n denoting the principle quantum number using the atomic convention
given later. Pn−1
The level degeneracy is = (2s + 1) `=1 (2` + 1) = 2n2 , including the spin degree of freedom.
This scheme for Hydrogen, when extended to Z ≥ 2 and with the Pauli principle forms the basis
for our understanding of the periodic table, atomic and indeed molecular physics, valence bonding
and Chemistry.
The various wave functions can be given in general in terms of the Laguerre L2`+1 n+` (ρ) polynomials
used last term. The result is:
ρ
ϕn`m` sms (~r) = −Nn` ρ` e− 2 L2`+1
n+` (ρ) Y`m` (r̂) | sms >, (10.5)
CHAPTER 10. QUANTUM DYNAMICS 261

where ρ = r/(An/2Z). Note that n appears in ρ and one needs to watch that aspect. The
normalization is given by
2 2Z 3 (n − ` − 1)!
Nn` =( ) .
An 2n[(n + `)!]3
For Hydrogen Z = 1. Here A = h̄2 /(µe2 ) = 0.5292 × 10−8 cm, is the Bohr radius which sets the
size (scale) of the atom. The Laguerre polynomial is defined by the generating function ( which is
required to stipulate all of the conventions) is
ρτ ∞
e− 1−τ X τ σ−a
Ga (ρ, τ ) = a+1
= (−1)a Laσ (ρ). (10.6)
(1 − τ ) σ!
σ≥a

Note Laσ (ρ) has σ − a zeroes.


One of the conventions used is that the radial wave function is positive at the origin r = 0. Once
chosen, the convention must be kept; otherwise off-diagonal matrix elements will be of the wrong
sign. Some of the most often used wave functions are listed below, and also drawn using this
convention. The Laguerre polynomials are given by the series:
σ−a
X (−1)λ+1 [σ!]2 ρλ
Laσ (ρ) = . (10.7)
(σ − a − λ)! (a + λ)! (λ)!
λ=0

For the case of σ = a + 1, this yields Laa+1 (ρ) = −(a + 1)! {a + 1 − ρ}, and for σ = a, Laa (ρ) = −(a)!.
Another important aspect of the Laguerre polynomials, which we note now for later use is:
−(σ!)2
Laσ (ρ = 0) = .
(σ − a)! a!
Note σ → n + ` and a → 2` + 1.
Using these properties, we can now write the various Hydrogen wave functions classified according
to the number of nodes in L2`+1 2`+1
n+` , which is n + ` − (2` + 1) = n − ` − 1 = number of nodes in Ln+` .
Thus we use the atomic convention for the principle quantum number n that n ≥ ` + 1, and
n − ` − 1 = number of nodes in the wave function, not counting the r = 0 or r = ∞ points.
The nodeless n = ` + 1 radial wave function is
Z 3 2
Rn` (r) = ( ) 2 p ρ`n e−ρn /2 .
A (` + 1)2 (2` + 1)!
The one node n = ` + 2 radial wave function is
Z 3 2
Rn` (r) = ( ) 2 p ρ`n (2` + 2 − ρn ) e−ρn /2 .
A 2
(` + 2) (2` + 2)!
The two node n = ` + 3 radial wave function is
Rn` (r) = · · · .
We shall use some of these wave functions in later problems.
Note again that ρ depends on n since ρ = r/(An/2Z).
Note that for the nodeless wave functions (1s, 2p, 3d, · · ·) the maximum moves out due to the
centrifugal barrier. Also the one node cases (2s, 3p, 4d, · · ·) have their one zero move out with
increasing ` value. Similar statements hold for the wave functions with more nodes.
These properties of the radial wave functions are included in the following expectation values for
the Hydrogen case
Z ∞
1 1 1 1 1
< n` | 3 | n` > = | Rn` (r) |2 r2 dr 3 = 3 3 (10.8)
r 0 r A n (` + 1)(` + 12 ) `
Z ∞
1 1 1
< n` | | n` > = | Rn` (r) |2 rdr =
r A n2
Z0 ∞
1 1 1 1
< n` | 2 | n` > = | Rn` (r) |2 dr = 2 3 1
r 0 A n (` + 2)
Note in the first result above there seems to be a problem at ` = 0, but that gets resolved by the
Dirac equation later. These results can be obtained by use of the generating function for the
Laguerre polynomials, we omit presenting this approach for simplicity.
CHAPTER 10. QUANTUM DYNAMICS 262

Zr
1s Z 3/2
R10 = ( A ) 2e− A

Zr
2s Z 3/2 1
R20 = ( A ) 23/2 (2 − Zr
A ) e− 2A

Zr
Z 3/2 √ − 2A
2p R21 = ( A ) 2 1 6 Zr
A e

4Z 2 r 2 Zr
3s Z 3/2 √
R30 = ( A ) 9 1 3 (6 − 4Zr
A + 9A2 ) e− 3A

Zr
3p R31 = ( A ) 9 1 6 2Zr
Z 3/2 √
3A (4 −
2Zr
3A ) e− 3A

Z 3/2 1 2Zr − 3A Zr
3d R32 = ( A ) 9 3A e

Table 10.1: Hydrogenic radial wave functions

un{ un{
1.4 1
1.2 0.75
1 ˜ u10 0.5 ˜ u21
0.8 0.25
0.6 r
0.4 -0.25 2 4 6 8
-0.5 ˜ u20
0.2
r -0.75
2 4 6 8 -1
un{
1
0.75 ‡u32
0.5
0.25 ˜ u30
r
-0.25 2.5 5 7.5 10 12.5 15
-0.5 ˆ u31
-0.75
-1
Figure 10.3: Hydrogen radial wave functions un` = rRn` .

10.3.2 Perturbations of the Hydrogen Atom


We can now consider two important internal corrections to the Hamiltonian. Later, external fields
will be discussed. Although we shall treat the following corrections using non-relativistic theory
(following Jordan & Heisenberg), the Dirac equation and Quantum Electrodynamics will provide
the complete theory of Hydrogen.
The perturbations are: first a correction for relativistic kinetic energy
p p2 1 p2 1
K.E. = p2 c2 + µ2 c4 − µc2 ' − ( )2 2 ,
2µ 2 2µ µc
2
p 2
VR = − 21 ( 2µ ) 1
µc2 .
CHAPTER 10. QUANTUM DYNAMICS 263

The second correction is due to the interaction of the magnetic moment of the electron with the
magnetic moment it “sees” as it rotates about the proton. To treat this effect correctly, one must
properly transform to the electron’s frame e This leads to the correction

1 1 2 ~ 1 dV 1 1 e2
VLS = ( ) ` · ~s = ( )2 3 (~` · ~s),
2 µc r dr 2 µc r
using V = −e2 /r. The above spin-orbit interaction correctly includes the relativistic effect of the
Thomas precession.
Therefore, we have the Hamiltonian for Hydrogen

p2 e2 1 p4 1 e ~` · ~s
H = H0 + VR + VLS = − − 3 2
+ ( )2 3 (10.9)
2µ r 8µ c 2 µc r

These corrections are small, of order α2 × ε0 , and hence first order perturbation theory is used.
The appearance of the ~` · ~s operator suggests we form eigenstates of j 2 , jz , `2 and s2 instead of our
original case of ϕn`m` ,sms .f we have
X
< ~r | ϕn`m` ,sms > = Rn` < jM | `m` sms > Y`m` (r̂) < ξ | χsms > (10.10)
m` ,ms
jm
= Rn` Y`s (r̂, ξ),

where a spin-space label ξ has been included for later use.


The first order correction then follows:
Z ∞
1 e 2 r2 dr
< n`jm | VLS | n`jm > = ( ) | Rn` |2 < ~` · ~s >
2 µc 0 r3
h̄2 ` 1
 
1 e 2 2 ; for j = ` + 2
= ( ) 1 `+1 1
2 µc A3 n3 (` + 1)(` + 2 ) ` − 2 ; for j = ` − 2
 1 1

1 eh̄ 2 1 1 1 `+1 ; for j = ` + 2
= ( ) . (10.11)
2 µc A3 n3 2` + 1 − 1` ; for j = ` − 12
2
e2 h̄2 2 µc2
This can be further simplified using µ2 c2 ( µe
h̄2
)3 = α2 α2 µc2 , and εn = − α
n2 2 , to

α2 1
; for j = ` + 12
 
− `+1
< n`jm | VLS | n`jm >= εn . (10.12)
n(2` + 1) + ` ; for j = ` − 12
1

The correction for the other term is


1 1
< n`jm | VR | n`jm >= − < n`jm | p4 | n`jm > . (10.13)
8 µ3 c2
e2
Since 2µ(H0 + r ) = p2 , this can be simplified

4 1 e2 2
< n`jm | VR | n`jm > = − < n`jm | (H0 + ) | n`jm >
8 µc2 r
1 1 1
= − { ε2n + 2εn e2 < > +e4 < 2 >}
2µc2 r r
εn n 3
= α2 2 { − }. (10.14)
n ` + 21 4

Combining these two corrections, we have

α2 n α2 1
; for j = ` + 12
 
3 − `+1
E =< n`jm | H0 +VR +VLS | n`jm >= εn {1+ 2 [ − ]+ }
n ` + 12 4 n(2` + 1) + 1` ; for j = ` − 21
(10.15)
e As done originally by A. Thomas.
f This is a small example of using j − j coupling instead of L − S coupling.
CHAPTER 10. QUANTUM DYNAMICS 264

which finally yields


α2 n 3
E = εn {1 + [ 1 − ]} . (10.16)
n2 j + 2
4
This is Sommerfeld’s result, which he obtained from different reasoning. Note that for the two
cases j = ` ± 21 , one gets that same result depending only on j, since
1
; for j = ` + 21
 
1 1 − `+1 1
+ =
`+ 1
2
2` + 1 + ` ; for j = ` − 21
1
j+ 1
2

Note that the level energy depends on n and j and therefore there is a degeneracy now on j, such
as for 2s1/2 , 2p1/2 and 3p3/2 , 3d3/2 .

10.3.3 Fine Structure


We found in the previous section that

α2 n 3
E = εn {1 + [ 1 − ]} , (10.17)
n2 j + 2
4
which results from including the spin-orbit and relativistic corrections in first order. The above
result
H was also obtained earlier by Sommerfeld, using the old Bohr model and the quantization rule
pdq = nh for each canonically conjugate pair of variables. He thus incorporated relativistic
effects and non-spherical orbital motion. There is however an important difference between the
present solution and the one provided by Sommerfeld. The difference is that the total number of
states, including degeneracy, is almost twice the number of states obtained from the Sommerfeld
theory. The increase in the total number of states of course arises from the introduction of electron
spin along with the rules for adding angular momentum. Although the same energy arises, in the
presence of external Electric and Magnetic fields the two theories differ, with the Schrödinger
theory yielding the experimental behavior of the spectrum ( and the line intensities as well).
Before considering the behavior of Hydrogen in external fields, let us consider the magnitudes of
the fine structure given above. One finds that the 1s, 2s and 2p levels are changed. We consider
the | n`jm > states
1s1/2 2s1/2 2p1/2 2p3/2 .
Since Enj has a j−degeneracy, the 2s1/2 and 2p1/2 have the same energy. One finds:
CHAPTER 10. QUANTUM DYNAMICS 265

2s, 2p (−3.4 ev)


H
HH 2p3/2 i − 0.113 × 10−4 ev
H
H i − 0.453 × 10−4 ev
2s1/2 , 2p1/2

1s (−13.605 ev)
H
HH
H i − 1.81 × 10−4 ev
H
1s1/2

Figure 10.4: Splitting of the n = 1 and n = 2 levels of Hydrogen due to relativistic and spin-
orbit splitting. Levels are not drawn to scale. The difference between the 2p3/2 and 2p1/2 levels is
0.453 × 10−4 e.v ≡ 10, 950 M hz. The 2p1/2 and s1/2 levels are still degenerate, but that degeneracy
is lifted by the Lamb shift.

Clearly the fine structure is small, but it is observable.


Another effect that one should consider is the finite size of the proton. However, since the proton’s
size is a ' 0.8 × 10−13 cm, which is quite small, one suspects that the proton’s size will cause a very
small change on the atomic spectrum, because a  A. The result found in the Homework for
Hydrogen confirms this remark, i.e.

∆(1s1/2 ) ' +12.5 × 10−10 ev


∆(2s1/2 ) ' +1.56 × 10−10 ev
∆(p) ' 0

For heavier nuclei, this effect increases and one must then include the finite nuclear size; in fact,
this is one way to learn about nuclear sizes, even isotopic shifts. For heavier particles in Bohr
orbit, such as µ−mesic, π − , K − anti-protonic atoms, etc., the finite nuclear size is appreciable and
is an important way of learning about the nuclear surface and strong interactions.
Although the proton size effect does remove the degeneracy of the Hydrogen 2s1/2 and 2p1/2 levels,
it does not explain the experimental result of a 2s1/2 − 2p1/2 shift of 4.38 × 10−6 ev, which is called
the Lamb shift (Lamb-Retherford 1950). g
Explanation of the Lamb shift is one of the great triumphs of quantum electrodynamics, for which
Feynman, Schwinger and Tomonoga won the Nobel prize in 1965. Such shifts occur in other states
and other atoms also.

10.3.4 The Zeeman and Paschen-Back Effects


Reference: Messiah p 706ff
Imposing external magnetic fields on the Hydrogen atom leads to changes in the spectrum. That
effect shows the electrical nature of atoms, but beyond that general observation one can explain
the details of the spectrum only after the proper number of states and their quantum numbers are
introduced. Here the Sommerfeld theory breaks down, and the Schrödinger equation including spin
triumphs!
g Willis Eugene Lamb Jr. received 1955 Nobel prize for discoveries regarding the hyperfine structure of hydrogen

spectrum.
CHAPTER 10. QUANTUM DYNAMICS 266

The splitting of spectral lines in external magnetic fields is an important method for detecting the
presence of magnetic fields in astrophysics. One detects shifts in the frequency of hydrogen lines,
for example, and can often attribute that shift in part to the imposed magnetic field.
The effect of an external magnetic field is represented by adding a term to the Hamiltonian; one
has the extra energy −~ µ·B ~ = −µz Bz ≡ µz B. (Here we ignore the quadratic B 2 term, which will
be examined later and seen to be related to diamagnetism) The magnet moment arise from both
orbital and spin angular momentum. Our Hamiltonian is now extended to:

H = H0 + H1 + HM
p2 e2
H0 = −
2µ r
1 p4 1 e ~` · ~s
H1 = HR + HLS = − 3 2
+ ( )2 3
8µ c 2 µc r
µ0 B
HM = (`z + 2sz ) (10.18)

The eigenstates for H0 are | n`m` sms >; whereas, for H0 + H1 , we use states | n`sjm > so as to
diagonalize the spin-orbit term. Note also that H is invariant under parity.
How shall we treat the magnetic term HM ? The treatment depends on the strength of the
magnetic field. For “weak” magnetic fields,

µ0 B  0.57 × 10−4 ev

(the fine structure of the 2s, 2p levels) we have the Zeeman effect. Since µ0 = 0.579 × 10−8 ev,
“weak” means B  104 gauss = 10 kilogauss.
For this situation, the magnetic energy is small compared to the fine structure correction, and thus
one treats HM as a perturbation of the Hamiltonian H0 + H1 . The basis states of H0 + H1
(namely, | `sjm > states) are used and we have for B  104 gauss
µ0 B
EB = Enj + < n`sjm | `z + 2sz | n`sjm >

µ0 B
= Enj + < n`sjm | jz + sz | n`sjm > . (10.19)

Here Enj can include all of the fine structure. Since, using C-G tables we have
X m
< n`sjm | sz | n`sjm >= |< jm | `m` sms >|2 h̄ms = ±
m m
2` +1
` s

for j = ` ± 1/2, we get


2` + 1 ± 1
EB = Enj + µ0 B[ ] j = ` ± 1/2 (10.20)
2` + 1
or
2j + 1
EB = Enj + µ0 B[ ], (10.21)
2` + 1
which is the Zeeman effect result.
Another way to evaluate the expression above, which is of use for other atoms, is to invoke the
Landé formula:

µ0 B
1 h̄ < n`sjm | [(~j + ~s) · ~j]jz | n`sjm >
< n`sjm | `z + 2sz | n`sjm > =
h̄ h̄2 j(j + 1)
m j(j + 1) − `(` + 1) + s(s + 1)
= [j(j + 1) +
j(j + 1) 2
= gJ m, (10.22)
j(j+1)−`(`+1)+s(s+1)
where the Landé g-factor is gJ = 1 + 2j(j+1) . With j = ` ± 1/2, this yields µ0 B[ 2j+1
2`+1 ]
as before.
CHAPTER 10. QUANTUM DYNAMICS 267

Note that non-degenerate perturbation theory suffices, since the 2s, 2p levels are not coupled by
HM . The states p1/2 , p3/2 can mix and then higher order terms are needed.
Thus we can write the Zeeman splitting as:

EB − En` = µ0 Bmgj ,

which can be generalized for Z ≥ 2 atoms. The important feature of this result is that it explains
the normal (or classical) Zeeman effect and the so-called anomalous Zeeman effect. The anomalous
Zeeman effect could not be explained by the classical picture of atomic currents; it had too many
lines. Now with the quantum theory and spin, the full splitting of spectral lines is obtained along
with the correct polarizations and intensities, which we will later explain.
The Zeeman splitting for the first few levels of hydrogen is displayed below.
insert figure here

Large Magnetic Fields Paschen-Back Limit


For large magnetic field B  104 gauss, the magnetic energy is larger than the fine structure
energy term H1 and thus H1 is the perturbation and we treat H0 + HM exactly. It is possible to
treat H0 + HM exactly because the eigenstates of H0 | n`m` sms > already diagonalize HM ;
hence , the Paschen-Back energy is

EB = εn + µ0 B(m` + 2ms )
= εn + µ0 B(m + ms )
µc2 1
εn = −α2 . (10.23)
2 n2
This Paschen-Back (PB) limit holds for large magnetic field and/or high n. Note that since the
spin-orbit interaction is small (∼ 1/n3 ) for large n, the PB limit applies to large n levels also-even
at lower magnetic field strengths.
Also note that the magnetic field is large here, but can not be so large as to introduce the
e2 2 e2 2 2 2
quadratic terms 2µc 2 A = 2µc2 r sin θB that have been neglected in HM . h
The first few levels in Hydrogen in the large magnetic field case are split as shown below: where
the m` ms values have been indicated as (m` ms ). Note that for n = 2, there are 6+2=8 states=
2n2 = total degeneracy, with two levels remaining at the value ε2 .

(1, +1/2) +2µ0 B

(0, +1/2) +µ0 B (0, +1/2) +µ0 B


 
 
2s  2p 
HH

H (1, −1/2); (−1, +1/2)
H @HH
HH @ HH (0, −1/2) −µ B
(0, −1/2) −µ0 B @ 0
@
(−1, −1/2)
@ −2µ B
0
(0, +1/2) +µ0 B


1s 
HH
H
HH
(0, −1/2) −µ0 B

Figure 10.5: Splitting of the n = 1 and n = 2 levels of Hydrogen due to a strong magnetic field; the
Paschen-Back limit. Levels are not drawn to scale. There are 8 levels for n = 2. Six of these are
split. Two of them (2p) for (m` , ms ) = (1, −1/2); (−1, +1/2) stay at the unperturbed energy ε2 .

h From the context, it should be clear when µ refers to the reduced mass and when it denotes the magnetic moment.
CHAPTER 10. QUANTUM DYNAMICS 268

Intermediate Magnetic Fields


For weak (B  104 gauss) and strong (B  104 gauss) magnetic fields, the splitting due to HM
has been discussed. The problem now is to determine the splitting fo the intermediate fields
(B ' 104 gauss. Since for fields of this strength the fine structure and magnetic terms H1 and HM
are comparable, we can not use perturbation theory and instead form the appropriate secular
determinant. Some of the symmetries help to simplify that problem to solving a 2 × 2 matrix
equation.
Our secular equation is, using the basis states | n`sjm >:

det Ω(E) = 0

with
Ω =< n`sjm | E − H0 − H1 − HM | n0 `0 sj 0 m0 > .
We now truncate this problem to n = 1,and 2 levels only; this step simplifies the problem to
dealing with the states 1s1/2 ; 2s1/2 , 2p1/2 ; 2p3/2 . Admixture of other states with n > 2 is possible
but requires considerable energy as is seen from the second order perturbation expression. The
parity operator commutes with H = H0 + H1 + HM and hence the ` = 0 and ` = 1 levels do not
mix; i.e. the matrix elements < s | H | p >= 0. Also the operator Jz commutes with H1 and HM ,
therefore states with different m-values do not mix; [Jz , H] = 0 ... < m | H | m0 >= 0, unless
m = m0 . Because of parity and cylindrical symmetry, the only states that mix (have off-diagonal
terms in Ω) are the | p1/2 m = ±1/2 > and states | p3/2 m = ±1/2 > .
The | s1/2 m = ±1/2 > and | s3/2 m = ±3/2 > energies are given by the original Zeeman result:

E(s1/2 ) = En + µ0 B2m = En0 ± µ0 B, (10.24)

for 1s1/2 and 2s1/2 levels, and


4
E(p3/2 ) = E2 + µ0 B m = E21 2 ± µ0 B, (10.25)
3
only for m = ±3/2.
For m = ±1/2 and | p1/2 m = ±1/2 > and | p3/2 m = ±1/2 >, the matrix is split into two separate
2 × 2 matrices- one for each value of m. We need to consider

Ωmm0 = (E − E2j ) δjj 0 − < n`jm | HM | n`j 0 m > (10.26)

where ` = 1 and the energy eigenstates of H0 + H1 are indicated by Enj for n = 2 and
j = 1/2, 3/2. We need the following matrix elements of HM = µ0h̄B (jz + sz )

2j + 1
< n`jm | HM | n`jm > = µ0 Bm
2` + 1
µ0 B
< n`jm | HM | n`j 0 m > = < n`jm | (jz + sz ) | n`j 0 m >

µ0 B
= < n`jm | sz | n`j 0 m >

µ0 B X
= < jm | `m` sms > ms < `ml sms | j 0 m >
h̄ m m
` s
q
( 2 + m)( 23 − m)
3
= −µ0 B
3
(10.27)

in the last line we need the C-G coefficients.


Using the above matrix elements, we can construct the secular determinant in the
m = ±1/2, p1/2 , p3/2 , space
q
E − E2 12 − µ0 Bm 23 1
µ0 B 94 − m2

q 3 = 0.
1
3 µ0 B 94 − m2 E − E2 32 − µ0 Bm 34

CHAPTER 10. QUANTUM DYNAMICS 269

The above represents two equations, one for m = +1/2, another for m = −1/2. it has the solutions
s
2 + ∆ 1 m2 1 ∆ m2 2
E± = + µ0 B{m ± ( − )+ ( + ) }, (10.28)
2 4 9 4 µ0 B 3

where  ≡ E2 12 , and ∆ ≡ E2 32 − E2 12 .

In the weak field limit ∆  µ0 B the above yields the Zeeman result and for strong fields ∆  µ0 B
the above yields the Paschen-Back result.
CHAPTER 10. QUANTUM DYNAMICS 270

insert a plot here.


Stark Effect–see HW and to be inserted here
CHAPTER 10. QUANTUM DYNAMICS 271

10.4 Time Dependent State Perturbation Theory


Reference: Messiah Chapter XVI.
The discussion of perturbation theory has so far been based on the stationary states or on
solutions of the time-independent Schrödinger equation. Now we turn to the time-dependent
Schrödinger equation and the corresponding treatment of time-dependent perturbation theory.
The motivation for discussing time-dependent perturbation theory is that one often deals with a
Hamiltonian which is subject to a time dependent interaction, i.e.

H = H0 + V (t). (10.29)

The meaning of V (t) is that energy is not conserved, but is instead leaving or entering the system.
This situation arises when the system ( say, an atom) is coupled to some external system. For
example, a time-dependent interaction occurs when two pendulums are coupled as shown.

 B
 B
 B
 BB
1m @ @ @
m
@ 2

Figure 10.6: The atomic system can be thought of pendulum 1 coupled to another pendulum 2
representing the electromagnetic field. Indeed, later the normal modes of the electromagnetic field
will be a set of pendula.

The Hamiltonian for system 1 ( the atom) is then time-dependent, since energy can be exchanged
with system 2 ( the electromagnetic field). It is this analogy that forms the basic idea of how
radiation from atoms occurs. The atomic system is coupled to the electromagnetic system, which
can receive and give energy to the atom. Thus, the atomic Hamiltonian would be time dependent.
Another example of how time dependent interactions occur in physical problems are the
mechanism of Coulomb excitation of nuclei and of atomic ionization and excitation. In Coulomb
excitation a heavy charged particle ( say an α particle) passes near a nucleus and interacts with the
nucleus through a Coulomb interaction V (t) = Ze2 / | ~r − R(t)~ |, where R(t)
~ describes the classical
orbit of the α particle. (One can use a classical description of the α particle trajectory because it is
heavy and of relatively low energy.) There results a nuclear excitation with energy pumped into
the system( nucleus) via V (t). The atomic case of a charged particle passing by the electron cloud
of atoms can also lead to excitation of electrons, or to ionization, which forms the basic mechanism
for particles leaving trajectories in various particle detectors such as the Wilson cloud chamber.
The main point is that a time-dependent interaction V (t) is encountered when energy can be fed in
to and out of a system. We must therefore consider a treatment involving the time development of
quantum states.
We seek to explain transitions, selection rules and line intensities and also rates of decay.

10.5 Time Evolution in Quantum Mechanics


The Schrödinger equation gives us the time development of the state vector | Ψ(t) > from which all
quantities of interest can be obtained

ih̄ | Ψ(t) >= H | Ψ(t) > . (10.30)
∂t
For a time-independent Hamiltonian, this equation can be written as:
i
| Ψ(t) >= e− h̄ H(t−t0 ) | Ψ(t0 ) >, (10.31)
∂ ∂
as seen by forming ih̄ ∂t | Ψ(t) >= H | Ψ(t) > + terms in ∂t H, which are zero. We have used this
form in building the Green’s function for free particle wave packets. Now we wish to generalize the
CHAPTER 10. QUANTUM DYNAMICS 272

above case to include Hamiltonians that depend on time. That generalization consists of taking
the Schrödinger equation as

ih̄ | Ψ(t) >= H(t) | Ψ(t) > . (10.32)
∂t
Another way to generalize the above time-dependent case to include H(t) is to take
i
| Ψ(t0 ) >= (1 − H(t)δt) | Ψ(t) >,

for δt = t0 − t infinitesimal. In fact, the two statements are equivalent since
| Ψ(t + δt) > − | Ψ(t) > ∂
ih̄ = ih̄ | Ψ(t) >= H(t) | Ψ(t) > . (10.33)
δt ∂t
The above steps are presented to show that one can readily incorporate a time dependent
Hamiltonian first in a differential form and secondly as an infinitesimal canonical transformation.
We have, in general, that
| Ψ(t) >= U (t | t0 ) | Ψ(t0 ) > (10.34)
which can be viewed as a transformation from LAB at t0 to LAB g at time t; i.e. a canonical
transformation. The above operator U is a unitary operator because the system has a fixed
normalization, i.e. the system doesn’t disappear. Energy might flow out in other forms, but we
still have the system in hand. In contrast, if the flux of material particles is nonzero, then U † U 6= 1
and we have a non-hermitian potential H † 6= H. One could introduce energy into the system
during the time interval t − t0 , and yet know that no flux is lost; i.e. the atomic system stays in
the laboratory and no flux leaves in the form of material particles from the atom. Thus although
time-dependent, the Hamiltonian is assumed to be Hermitian and therefore

< Ψ(t) | Ψ(t) >=< Ψ(t) | U † (t | t0 )U (t | t0 ) | Ψ(t) >=< Ψ(t0 ) | Ψ(t0 ) >= 1, (10.35)

or we have a unitary transformation

U † (t | t0 )U (t | t0 ) = 1. (10.36)

The inverse operator is clearly defined by U −1 (t | t0 ) = U (t0 | t) and hence

U † (t | t0 ) = U −1 (t | t0 ) = U (t0 | t). (10.37)

Finally, the time evolution operator U has the continuous group properties

U (t0 | t0 ) = 1
0
U (t | t) U (t | t0 ) = U (t0 | t0 ) , (10.38)

where the last relationship says that two time displacements can be equivalent to one
displacement, i.e. we have the group property. If the second time displacement is infinitesimal,
then the above group property becomes with δt = t0 − t
i
[1 − H(t)δt]U (t | t0 ) = U (t + δt | t0 )

or
U (t + δt | t0 ) − U (t | t0 )
H(t)U (t | t0 ) = ih̄
δt

H(t)U (t | t0 ) = ih̄ U (t | t0 ) . (10.39)
∂t
Another way to obtain this result is by taking
∂ ∂
ih̄ | Ψ(t) > = ih̄ U (t | t0 ) | Ψ(t0 ) >
∂t ∂t
= H(t) U (t | t0 ) | Ψ(t0 ) > (10.40)

or

ih̄ U (t | t0 ) = H(t) U (t | t0 ).
∂t
CHAPTER 10. QUANTUM DYNAMICS 273

If H(t) is independent of time H(t) → H, and the solution of the above operator equation is
i
U (t | t0 ) = e− h̄ H(t−t0 ) = U † (t0 | t) 7−→ 1,
t→t0

all for a time independent H. In general, one can not write U (t | t0 ) in this simple form; instead,
the integral equation form serves as a starting point for finding U (t | t0 ) when H(t) occurs
Z t
i
U (t | t0 ) = 1 − H(τ ) U (τ | t0 )dτ. (10.41)
h̄ t0

10.5.1 Schrödinger and Heisenberg Pictures


The time evolution operator U (t | t0 ) can be determined from the above integral equation once
H(t) is known. Another feature of U (t | t0 ) is that it permits us to change our description of
dynamics. One can take the view that the state vectors rotate in Hilbert space with time and that
the operators ( aside from questions of explicit time dependence) are fixed. This picture is called
the Schrödinger picture. In the Schrödinger picture, the viewpoint taken is of a time-dependent
state vector
| ΨS (t) >= U (t | t0 ) | ΨS (t0 ) >,
where | ΨS (t) > rotates but | ΨS (t0 ) > is a fixed vector in Hilbert space. The Schrödinger picture
~ ~`, ~σ , p2 + V (r), etc. ( Here we put aside for
also uses time-independent operators, such as p~ = h̄i ∇, 2m
a while, the possibility of some explicit time dependence in the Schrödinger operators.) This
“picture” is the one we have been using all along for which we calculate probabilities in the
following manner. The expectation value for an operator ΩS is
X
< ΩS >t =< ΨS (t) | ΩS | ΨS (t) >= ωn |< n | ΨS (t) >|2 , (10.42)
n

where the subscript S denotes operators and states in the Schrödinger picture and < ΩS >t
denotes the expectation value. The operator ΩS denotes an observable with eigenvalues
ΩS | n >= ωn | n > . Correspondingly, the probability for finding the system described by
| ΨS (t) > with the eigenvalue ωn at the time t is

pn (t) =|< n | ΨS (t) >|2 =|< n | U (t | t0 ) | ΨS (t0 ) >|2 =| cn (t) |2 . (10.43)

If at the initial time ( in the far past t0 → −∞) the system is known to be in a state n0 then at
some later time the probability of finding it in | n > is

pnn0 (t) =|< n | U (t | −∞) | n0 >|2 =| cnn0 (t) |2 . (10.44)

For large time lapses t → +∞, this equation becomes

pnn0 =|< n | U (+∞ | −∞) | n0 >|2 =|< n | S | n0 >|2 , (10.45)

which defines the so-called S-matrix, which was introduced originally by Wheeler and later by
Heisenberg. Note cnn0 (+∞) ≡ Snn0 .
The probability is properly normalized since
X
pn (t) =< Ψ(t0 ) | U † (t | t0 ) U (t | t0 ) | Ψ(t0 ) >=< Ψ(t0 ) | Ψ(t0 ) >= 1, (10.46)
n

which verifies that U is a unitary operator. We see that knowledge of the time evolution operator
is all we need to calculate expectation values and probabilities for transitions cnn0 . The goal
therefore is to develop methods for calculating U (t | t0 ).
In the Heisenberg picture another view is taken; namely, that the operators depend on time while
the state vectors are fixed. This picture is completely equivalent to the Schrödinger picture and is
simply related to it by a unitary transformation. Historically, the Heisenberg matrix mechanics
methods were developed first and then soon thereafter the Schrödinger wave mechanics appeared.
CHAPTER 10. QUANTUM DYNAMICS 274

Here we see that these two approaches are simply equivalent descriptions which are related by the
unitary operator U (t | t0 )
| ΨS (t) >= U (t | t0 ) | ΨH >, (10.47)
where | ΨH > is a fixed vector equal to the fixed vector | ΨS (t0 ) > . The “transformation theory”
is here expressed by the existence of a unitary transformation relating these two pictures of
quantum mechanics.
The equivalence of the two pictures is seen by introduction of the following “Heisenberg operators”
ΩH and the expectation value
< Ω >t =< ΨS (t) | ΩS | ΨS (t) >=< ΨH | ΩH (t) | ΨH >, (10.48)
where ΩS is independent of time aside from possible explicit time dependence. Here
ΩH (t) = U † (t | t0 ) ΩS U (t | t0 ). (10.49)
The dependence of < Ω >t on t is also given by
∂ ∂
ih̄ < Ω >t = < ΨS (t) | [ΩS , H] + ih̄ ΩS | ΨS (t) >
∂t ∂t
d
= < ΨS (t) | ih̄“
ΩS ” | ΨS (t) > (10.50)
dt
where we have included a term for a possible explicit time dependence in ΩS . This yields our old
result
d 1 ∂
“ ΩS ” = [ΩS , H] + ΩS ,
dt ih̄ ∂t
d d
which defines an operator “ dt ΩS ” whose expectation value constructs the derivative dt < Ω >t .
Now the Heisenberg picture describes the corresponding time dependence of < Ω >t by
d d
< Ω >t = < ΨH | ΩH (t) | ΨH >
dt dt
d 1 ∂
ΩH (t) = U † { [ΩS , H] + ΩS }U
dt ih̄ ∂t
1 ∂
= [ΩH (t), HH (t)] + U † (t | t0 ){ ΩS (t)}U (t | t0 )
ih̄ ∂t
1 ∂
= [ΩH (t), HH (t)] + ΩH (t). (10.51)
ih̄ ∂t
This equation is called the Heisenberg equation of motion, which can be used instead of the
Schrödinger equation to stipulate the time evolution of quantum mechanics. It has the main
advantage of dealing directly with the observables Ω. For example, with Ω → rH (t), or pH (t), these
equations are
d 1 d 1
rH (t) = [rH (t), HH (t)] and pH (t) = [pH (t), HH (t)] (10.52)
dt ih̄ dt ih̄
p2S
Note that HH (t) = U † (t | t0 )H(t)U (t | t0 ). If we have H(t) = 2m + V (rS ), then
p2H
HH (t) = 2m+ V (rH ), with rH , pH depending on time and satisfying the Heisenberg equations of
motion above.
If ΩS = H(t) is taken, the above results tell us that
d ∂ ∂
“ H(t)” = [H(t), H(t)] + H(t) = V (t) 6= 0,
dt ∂t ∂t
in general, or
d
< H(t) >6= 0,
dt
for a time-dependent interaction V (t), which shows that a time-dependent interaction corresponds
to non-conservation of energy for the system. Of course, the missing or gained energy is conserved
overall; it is transferred to another system, as in an atom absorbing or emitting electromagnetic
radiation. The hermiticity of the Hamiltonian is a statement that one still has all of the system in
hand, albeit with a different energy. It is the time-dependence that incorporates energy alterations
of the system (the atom).
CHAPTER 10. QUANTUM DYNAMICS 275

10.5.2 Dirac or Interaction Picture


The previous section showed that two pictures, related by a unitary operator U (t | t0 ), can be used
to describe quantum dynamics. Also, we saw that once U (t | t0 ) is known the probabilities and
associated expectation values could be obtained.
A third picture has been developed called the Dirac or Interaction picture, which is particularly
convenient for evaluating the time development operator. The basic idea is to separate out the
motion associated with H0 for which the eigenstates are supposedly known and thereby isolate the
time-dependent interaction term V (t) for special treatment. (Hence the origin of the term
“interaction picture.”) i
The basic idea then is to isolate the time dependence associated with H0 by taking
i
| ΨS (t) >= e− h̄ H0 t | ΨD (t) >, (10.53)
where | ΨD (t) > is the Dirac state vector, which depends on time since the total Hamiltonian and
not just H0 determines the full time evolution of | ΨS (t) > .
In the case that V (t) → 0, one has | ΨD (t) >→ a fixed state vector | ΨD > . With V (t) 6= 0, the
state | ΨD (t) > satisfies
∂ ∂
i
ih̄ | ΨS (t) > = H0 | ΨS (t) > +e− h̄ H0 t ih̄ | ΨD (t) >
∂t ∂t
= H0 | ΨS (t) > +V (t) | ΨS (t) > (10.54)
or

| ΨD (t) >= VD (t) | ΨD (t) >,
ih̄ (10.55)
∂t
with the Dirac operator defined by
i i
VD (t) ≡ e+ h̄ H0 t V (t)e− h̄ H0 t .
The evolution of | ΨD (t) > is determined by a unitary operator UD (t | t0 ), with all of the
aforementioned properties of U, except it is controlled by VD (t). We have
| ΨD (t) >= UD (t | t0 ) | ΨD (t0 ) > (10.56)
and from the above

ih̄UD (t | t0 ) = VD (t) UD (t | t0 ). (10.57)
∂t
Clearly UD (t0 | t0 ) = 1 and with that initial value we have the integral equation
i t
Z
UD (t | t0 ) = 1 − dτ VD (τ ) UD (τ | t0 ). (10.58)
h̄ t0
For no interaction VD (t) → 0, we get UD → 1, and | ΨD (t0 ) > is again a fixed vector.
The choice made above to isolate the effect of V (t) is available to us since no change occurs in
expectation values or probabilities; hence, the Dirac picture is just another canonical or unitary
transformation representing a change in description with no change in Physics. To see this point
consider the expectation values again
< Ω >t =< ΨS (t) | ΩS | ΨS (t) >=< ΨD (t) | ΩD (t) | ΨD (t) >, (10.59)
where the Dirac operators are defined by
i i
ΩD (t) = e+ h̄ H0 t ΩS e− h̄ H0 t .
Correspondingly, the time development of ΩD (t) is controlled by H0 since
d 1
ΩD (t) = [ΩD (t), H0 ].
dt ih̄
Thus in this discussion the operators move according to H0 , whereas the states | ΨD (t) > move
according to VD (t). This rendition is the Dirac picture in “Schrödinger style,” also know as the free
operator description
i Everyone, except Dirac, also calls this the “Dirac picture.” Ironically, Dirac indicated in his famous 1963-1964

Belfer lectures that this picture might not be valid for generalized non-holonomic mechanics, which apparently plays
a role in quantization in curved space in General Relativity. He also indicated that introduction of the Dirac picture
might be behind the singularity difficulties of QED.
CHAPTER 10. QUANTUM DYNAMICS 276

Probabilities
We have shown that the probability for finding a state initially (at time t0 ) given by | Ψ(t0 ) > in
an eigenstate of Ω (| n >) at time t is

pn (t) =|< n | ΨS >|2 =|< n | U (t | t0 ) | Ψ(t0 ) >|2 . (10.60)

If | Ψ(t0 ) >→| n0 > at t0 → −∞, this becomes

pnn0 (t) =|< n | U (t | −∞) | n0 >|2 . (10.61)


i i
Now we can show that j U (t | t0 ) = e− h̄ H0 t UD (t | t0 ) e+ h̄ H0 t0 , hence
i
pn (t) = |< n | Ψs >|2 =|< n | UD (t | t0 )e+ h̄ H0 t0 | Ψ(t0 ) >|2
pnn0 (t) = |< n | UD (t | −∞) | n0 >|2 . (10.62)
0
The fact that | e+iεn t e−iεn t0 |= 1 has been used here.
This establishes that knowledge of UD (t | t0 ) is sufficient for determining all probabilities and
expectation values.

10.5.3 The Dyson Series


The determination of UD (t | t0 ) is needed to obtain transition probabilities. Clearly VD (t)
determines UD since
i t
Z
UD (t | t0 ) = 1 − dτ VD (τ ) UD (τ | t0 ). (10.63)
h̄ t0
Usually one hopes that the above equation can be iterated k and leads to a convergent series. For
strong couplings V (t) and/or very intense radiation fields, ( as in lasers) the convergence problem
is especially difficult. l
If one iterates the above integral equation, one gets

1 t
Z Z t Z τ
1
UD (t | t0 ) = 1 + dτ VD (τ ) + ( )2 dτ VD (τ ) dτ 0 VD (τ 0 ) + · · ·
ih̄ t0 ih̄ t0 t0
∞ Z t Z τ1 Z τs−1
X 1 s
= ( ) dτ1 VD (τ1 ) dτ2 VD (τ2 ) · · · dτs VD (τs ) . (10.64)
s=0
ih̄ t0 t0 t0

Note that the upper limits are intertwined, whereas a common lower limit t0 occurs. Let me state
without proof (see later) that the above result can be rewritten in terms of a fixed set of upper and
lower limits as the Dyson series:
∞ Z tZ t
X 1 1 s
UD (t | t0 ) = ( ) dτ1 dτ2 · · · dτs T [VD (τ1 )VD (τ2 ) · · · VD (τs )]
s=0
s! ih̄ t0 t0
Rt
− h̄i dτ VD (τ )
= T [e t0
], (10.65)

where T denotes a time-ordered product of the operators; for example for two operators:

VD (τ )VD (τ 0 ) for τ ≥ τ 0
T [VD (τ )VD (τ 0 )] = { (10.66)
VD (τ 0 )VD (τ ) for τ 0 ≥ τ

At first, we shall deal with only s ≤ 2 terms.


i i i
j Note we use: | ΨS (t) >= e− h̄ H0 t UD (t | t0 ) | ΨD (t0 ) >= e− h̄ H0 t UD (t | t0 ) e+ h̄ H0 t0 | ΨS (t0 ) >
= U (t | t0 ) | ΨS (t0 ) > .
k Or, to solve this equation, one might resort to non-iterative approaches, such as setting up a numerical grid of τ

values and/or assuming forms for the RHS and LHS, etc.
l For electrodynamics V (t) is made stronger by virtue of the long range nature of the Coulomb interaction, or the
D
zero mass of the photon, and by the occurrence of pair production processes.
CHAPTER 10. QUANTUM DYNAMICS 277

Proof of the Dyson Series


One way to prove the Dyson series and to define time-ordered products for more than two
interactions is to divide the integration into zones and to reorder the integration steps. That
approach will not be presented here. Instead, a proof using the step function will be presented.
Recall that the step function is
1 for x ≥ 1
Θ(x) = { (10.67)
0 for x < 1
Consider the s = 2 term in the original series and introduce the step function
Z t Z τ1 Z t Z t
dτ1 VD (τ1 ) dτ2 VD (τ2 ) = dτ1 dτ2 VD (τ1 ) VD (τ2 ) Θ(τ1 − τ2 ). (10.68)
t0 t0 t0 t0

Using dummy variables, the above is also

1 t t
Z Z
= dτ1 dτ2 [VD (τ1 ) VD (τ2 ) Θ(τ1 − τ2 ) + VD (τ2 ) VD (τ1 ) Θ(τ2 − τ1 )]. (10.69)
2 t0 t0

Thus we have
Z t Z τ1 Z tZ t
1
dτ1 VD (τ1 ) dτ2 VD (τ2 ) = dτ1 dτ2 T [VD (τ1 ) VD (τ2 )], (10.70)
t0 t0 2! t0 t0

where the time ordered product for two potentials is defined by

T [VD (τ1 ) VD (τ2 )] ≡ VD (τ1 ) VD (τ2 ) Θ(τ1 − τ2 ) + VD (τ2 ) VD (τ1 ) Θ(τ2 − τ1 ), (10.71)

which is equivalent to our earlier expression.


This process continues for high terms. Carrying out the s = 3 case will suffice to show the general
result. For the s = 3 term of the original series, we again introduce step functions
Z t Z τ1 Z τ2 Z t Z t Z t
dτ1 VD (τ1 ) dτ2 VD (τ2 ) dτ3 VD (τ3 ) = dτ1 dτ2 dτ3 VD (τ1 ) VD (τ2 ) VD (τ3 ) Θ(τ1 −τ2 )Θ(τ2 −τ3 ).
t0 t0 t0 t0 t0 t0
(10.72)
Introducing dummy variables again, in this case we have 3! = 6 possible choices m
Z t Z τ1 Z τ2
1 t t
Z Z Z t
dτ1 VD (τ1 ) dτ2 VD (τ2 ) dτ3 VD (τ3 ) = dτ1 dτ2 dτ3 {
t0 t0 t0 3! t0 t0 t0
+ VD (τ1 ) VD (τ2 ) VD (τ3 ) Θ(τ1 − τ2 )Θ(τ2 − τ3 )
+ VD (τ1 ) VD (τ3 ) VD (τ2 ) Θ(τ1 − τ3 )Θ(τ3 − τ2 )
+ VD (τ2 ) VD (τ1 ) VD (τ3 ) Θ(τ2 − τ1 )Θ(τ1 − τ3 )
+ VD (τ2 ) VD (τ3 ) VD (τ1 ) Θ(τ2 − τ3 )Θ(τ3 − τ1 )
+ VD (τ3 ) VD (τ1 ) VD (τ2 ) Θ(τ3 − τ1 )Θ(τ1 − τ2 )
+ VD (τ3 ) VD (τ2 ) VD (τ1 ) Θ(τ3 − τ2 )Θ(τ2 − τ1 ) }
(10.73)

Thus we have the RHS is: Z tZ t


1
= dτ1 dτ2 T [VD (τ1 ) VD (τ2 )], (10.74)
3! t0 t0

where the time ordered product for three potentials is defined by


X
T [VD (τ1 ) VD (τ2 )VD (τ3 ) ] ≡ VD (τi ) VD (τj ) VD (τk ) Θ(τi − τj )Θ(τj − τk ), (10.75)
permute(i,j,k)

where the sum extends over the 3! permutations of 123.


m These are: 123; 132; 213; 231; 312; 321.
CHAPTER 10. QUANTUM DYNAMICS 278

The steps for higher terms should now be obvious, so we have for the sth term:
s Z
!
1 Y t
dτi T [VD (τ1 ) VD (τ2 ) · · · VD (τs )] (10.76)
s! 1 t0

where the time-ordered product is


X
T [VD (τ1 ) VD (τ2 ) · · · VD (τs ) ] ≡ VD (τi ) VD (τj ) · · · VD (τk ) Θ(τi − τj ) · · · Θ(τk−1 − τk ),
permute(1,2,···s)
(10.77)
where the sum extends over the s! permutations of 12 · · · s.
Our final result is that:
∞ Z tZ t
X 1 1 s
UD (t | t0 ) = ( ) dτ1 dτ2 · · · dτs T [VD (τ1 )VD (τ2 ) · · · VD (τs )]
s=0
s! ih̄ t0 t0
Rt
− h̄i dτ VD (τ )
= T [e t0
], (10.78)
n
where T denotes a time-ordered product of the operators.

Coupled Equations for the Probability Amplitude


If we introduce the definition pnn0 (t) =| cnn0 (t) |2 , with cnn0 (t) ≡< n | UD (t | −∞) | n0 >, the above
operator equations become very explicit coupled equations; namely, in integral equation form

1 X t
Z
cnn0 (t) = δnn0 + dτ < n | VD (τ ) | n00 > cn00 n0 (τ ) (10.79)
ih̄ 00 −∞
n

or as a differential equation
∂ X
ih̄ cnn0 (t) = eiωnn” t < n | Vt ) | n00 > cn00 n0 (τ ), (10.80)
∂t 00
n

where ωnn” = (εn − εn00 )/h̄ is the Bohr transition frequencies.


The iterated series for the n0 → n transition or probability amplitude, cnn0 , is then

1 t
Z X
cnn0 (t) = δnn0 + dτ1 < n | VD (τ1 ) | n0 >
ih̄ −∞
n0
Z t Z τ1
1
+ ( )2 dτ1 < n | VD (τ1 ) | n0 > dτ2 < n0 | VD (τ2 ) | n0 > + · · · (10.81)
ih̄ −∞ −∞

where < n | VD (τ ) | n0 >= eiωnn0 t < n | V (τ ) | n0 > .


The probability amplitude can also be used to construct the states:
i i
X X
| ΨS (t) >= | n >< n | UD (t | t0 ) | ΨD (t0 ) > e− h̄ εn t = | n > e− h̄ εn t cnn0 (t), (10.82)
n n

where we set | ΨD (t0 ) >→| n0 >, as t0 → −∞.

10.5.4 Perturbation theory–“Weak” V (t).


So far everything has been exact ( provided the Dyson series includes all terms). Now we consider
time-independent perturbation theory. The basic assumption is that V (t) is weak in some sense; it
may vary slowly or rapidly with time, but now we assume only that V (t) is weak. If V (t) is very
˙ 0 ≈= 0 or cnn0 ≈ δnn0 . This means that there is then little depletion of the initial
small, then ih̄cnn
state | n0 > . Mostly, the particle stays in state | n0 > with only a small flux flow into | n > .
n In the relativistic version the interaction is replaced by a scalar Lagrangian density L and the Dyson form looks
R t
i
− h̄ d4 x L(x)
like: T [e t0 ] where x is a space-time four vector.
CHAPTER 10. QUANTUM DYNAMICS 279

Later, it will be shown how Weiskopf and Wigner improved this assumption by accounting for the
depletion of the initial level | n0 > .
Physically, to satisfy cnn0 ≈ δnn0 , the transfer of energy must be weak or slow compared to the
electron’s motion. For example, an electron orbits many times before it emits radiation; in terms of
the period of orbital motion Torbit and the period associated with the decay time Tdecay , we have
Torbit << Tdecay , which means many orbits before decay occurs. Optical transitions in atoms take
about Tdecay ≈ 10−8 sec, whereas from the Bohr model we get Torbit ≈ 10−15 sec. Thus
Tdecay ≈ 107 Torbit , and the condition Torbit << Tdecay applies. So we have some basis for
considering the coupling of the atom to radiation as “weak.”
Now we iterate the equation for the probability amplitude cnn0 (t), and solve using various types of
V (t) which are of physical interest.

Various Cases for V (t).


Several types of time-dependence occur in physical situations.
• Constant for a time T .
For example, V (t) = V, for 0 ≤ t ≤ T, but is otherwise zero. This situation arises in decay
and for scattering problems. We will treat atomic radiation using this case.

• Periodic or HarmonicR an harmonic case is for example, V (t) = vω sin ωt. A periodic case
for example has V (t) = dωvω eiωt = V (t + T ). This situation occurs in studies of resonances
and decays, also for external driving fields.
• Sudden This is the case when V (t) turns on or on and off rapidly compared to the time of
motion. Examples of sudden cases occur when (a) electric or magnetic fields are suddenly
switched in direction (b) or fields are turned on/off suddenly, (c) The Coulomb interaction
Ze2 /r is suddenly changed when Z is changed by a decay such as β-decay.
• Adiabatic This term is used for cases where the interaction is turned on or off very slowly
compared to the times involved in the system’s motion. As a result of slowly turning on
V (t), the states of the system evolve slowly from the initial state to final states without
transient effects. See Messiah’s discussion of adiabatic turning on and the adiabatic theorem,
which states that if a system is in a state | n > and an adiabatic turning on and off occurs,
then the system returns to the state | n > at the end.

Constant V
The case of V (t) = V, for 0 ≤ t ≤ T, but is otherwise zero is of special interest. we can treat this
case explicitly as follows. For states n 6= n0 ,
Z T
1
cnn0 (t) = < n | V | n0 > eiωnn0 τ dτ + · · · (10.83)
ih̄ 0

Defining,
t
eiωT − 1
Z
f (ω, t) ≡ eiωτ dτ = ,
0 iω
we see that
1
cnn0 (t) = < n | V | n0 > f (ωnn0 , T ) + · · · (10.84)
ih̄
and
1
pnn0 (t) =| cnn0 (t) |2 = |< n | V | n0 >|2 | f (ωnn0 , T ) |2 , (10.85)
h̄2
is the probability of funding the system the state | n > if it was initially in the state | n0 > . Now
since f (ω, t) = eiωt/2 sin(ωt/2)/ ω2 we have

sin2 ( ω2 t) 2
| f (ω, t) |2 = t .
( ω2 t)2
CHAPTER 10. QUANTUM DYNAMICS 280

Plotting | f (ω, t) |2 versus ω one sees the usual shape


insert figure.
As time increases, the width ≈ 2π 2
t 2 decreases, while the height increases as t . A rough estimate of
2π 2 1
the area under the curve is then t 2 × t × 2 ≈ 2πt. More rigorously we have
Z ∞
dω | f (ω, t) |2 = 2πt,
−∞

or we can say that as t → +∞, | f (ω, t) |2 → 2πtδ(ω).


The above property shows that to get a physically meaningful quantity we should consider the
probability per time ( or the derivative dpnn0 /dt). At a give time we form,

dpnn0 1 1 | f (ωnn0 , t) |2
= | cnn0 (t) |2 = 2 |< n | V | n0 >|2 , (10.86)
dt t h̄ t
|f (ω
nn0 ,t)|2
where the limit t → ∞, can be taken at a later stage. Note that t 7−→ δ(ωn − ωn0 ), which
t→∞
is a statement of energy conservation for the system for this type of interaction.
The above result still yields a singularity, which corresponds to conservation of energy for the
whole system. To extract a physically meaningful (finite) transition rate ṗ, we must recognize that
transitions occur to levels that have a spread in energy. For example, a transition n0 → nf , can be
described as nf
}∆ωf


n0

Figure 10.7: The transition from level n0 to a finite set of levels with a spread or width ∆ωf .

The spread could be a set of near degenerate levels or due to a natural spread in energy for the
following reasons:
1. The atom is coupled to a continuum, the electromagnetic field, which can receive or input a
spread of energies to the particles
2. A natural line width exists (see item above) which can be assigned to each level
3. Measurement of the energy of the final state and the transition is limited experimentally by
the resolution of the apparatus for detecting photons.

4. A conceptual limitation on our resolution exists because for a finite time of measurement, an
uncertainty of energy, or our ability to measure energy, is introduced by the uncertainty
principle.
Here we visualize a coupled system of atom plus radiation and the states are real a product
| n >=| atom >| radiation >=| natom , nE&M >, i.e. we have two system coupled by V (t).
Conservation of energy applies to the whole system, but a transfer of energy can occur between the
atom and the electromagnetic system. Conservation of energy will later take the form of a
0
δ(ωatom + ωγ − ωatom − ωγ0 ), which here provides a way for fluctuation in photon (γ )energies to
cause a spread. For this constant potential example, no net energy is fed in or leaves the
subsystem –the atom. This will be be made clearer in the following section.
At this stage, we are not summing over final states that might be degenerate, with quantum
numbers, or attributes that we are not selecting out in our measurement process. For example, if a
spin-up and a spin down state have the same final energy and we count both in our detectors, we
should include a sum over spin of the final states. This will be explored later.
CHAPTER 10. QUANTUM DYNAMICS 281

Our procedure is therefore to average over the spread in the final levels energy spread and obtain
2
probability 1 X 2 | f (ωnn0 , t) |
< >averaged over ∆ωf ≡ Wn0 →nf = |< n | V | n 0 >| (10.87)
sec h̄2 nf t

Using box normalization for the final states (here we have some underlying continuum such as the
electromagnetic fields) Z 3
X Z d nf
Z
3
= d nf = dE = ρf (ωf )dωf .
n
dE
f

3 R∞
d n
where ρf (ωf ) ≡ dEf = the density of final states. The integral over 0 dωf is from zero to
infinity, but since the peak in f (ωnf − ωn0 , t) occurs at ωnf = ωn0 , we can extend the integral from
−∞ to +∞.
If both ρf and |< nf | V | n0 >|2 are smooth functions of the final energy ωf then in view of the
sharp peak in | f |2 /t, we can write

Wn0 →nf = |< nf | V | n0 >|2 ρf (ωn0 ) (10.88)

which is the probability/sec. Here, we evaluate the above expression with ωnf = ωn0 and apply it
for a particular final level nf . This is “Fermi’s Golden Rule.” We evaluate the above expression
with ωnf = ωn0 and apply it for a particular final level nf . Other transitions to other levels
n0f , n00f · · · can occur and often one sums over these levels as well, but that step involves discrete
quantum numbers. ( The rule is to average over initial states and sum over final discrete numbers,
if not observed.)
This discussion might be vague, because it is so general, but a better understanding of the
meaning of the above rule comes about by applying it to various situations. We shall therefore
consider Elastic Scattering as a means of gaining that understanding.

Elastic Scattering
An important example of a time-dependent interaction that can be described as being on for a
time, occurs in the scattering of a particle from a potential. The particle “sees” a potential
V (t) = V, for 0 ≤ t ≤ T, but is otherwise zero, during its passage across the scattering center.
Although we could consider the case of inelastic scattering. wherein energy is transferred to the
target and/or as radiation, we shall not discuss that case now.
The particle states are in the continuum, which using box normalization are:
~ ~
eik0 ·~r 2π 3 eik0 ·~r 2π 3
< ~r | n0 > = 3 =( )2 = ( ) 2 < ~r | ~k0 >
(L) 2 L (2π) 32 L
~ ~
eikf ·~r 2π 3 eikf ·~r 2π 3
< ~r | nf > = 3 =( )2 = ( ) 2 < ~r | ~kf > (10.89)
(L) 2 L (2π) 32 L

In the box. of volume L3 . the values of ~k are restricted by periodic boundary conditions to values
~k = 2π ~n, where ~n denotes nx , ny , nz which are integers.
L
Hence
L L
d3 n = ( )3 d3 k = ( )3 k 2 dkdΩk̂ (10.90)
2π 2π
equal the number of states. The density of final states is then

d3 n L dk
ρf = = ( )3 k 2 dΩk̂ . (10.91)
dE 2π dE
h̄2 kf2 dE h̄2
with E = 2m , dk m kf . and the density of states becomes

m L 3
ρf = ( ) kf dΩkˆf . (10.92)
h̄2 2π
CHAPTER 10. QUANTUM DYNAMICS 282

Using Fermi’s golden rule, we get

2π m L 3 (2π)6
Wk~0 →k~f = ( ) k f dΩ ˆ
kf |< ~kf | V | ~k0 >|2 , (10.93)
h̄ h̄2 2π L6
where we used
~ ~
e−ikf ·~r eik0 ·~r
Z
2π 3 2π 3 ~
< nf | V | n0 >= ( ) 3 V (r) 3 d3 r = ( ) < kf | V | ~k0 > .
L (2π) 2 (2π) 2 L

The transition rate above tells us the probability per second for scattering into a solid angle dΩkˆf .
h̄ 1
To get a cross section that flux must be divided by the incident flux, which is jincident = m k0 L3 ,
then
particles/sec into dΩkˆf
dσ =
incident flux
(2π)4 m
L3 h̄3 kf|< ~kf | V | ~k0 >|2 dΩkˆf
= h̄ 1
. (10.94)
m k0 L3

We find that

=| fBorn |2
dΩkˆf
where
m
fBorn = ±(2π)2 < ~kf | V | ~k0 >, (10.95)
h̄2
which is our old result for elastic scattering in the Born approximation. Higher order terms of the
time dependent theory can be evaluated and shown to yield the T-matrix of stationary state
scattering theory.
CHAPTER 10. QUANTUM DYNAMICS 283

10.6 The Free Electromagnetic Field


Let us consider an electromagnetic field without currents or sources, i.e. without the atomic
system. This case involves solving Maxwell’s equations o for the free electromagnetic fields. We
have:

~ ×B
~ 1 ∂ ~
∇ = E
c ∂t
~ ×E
~ 1 ∂ ~
∇ = − B
c ∂t
~ ·B
∇ ~ = 0
~
∇·E~ = 0 (10.96)
~ Φ,
Then we introduce the vector and scalar potentials A,
~ =∇
B ~ ×A
~

~ = −1 ∂ A
E ~ − ∇Φ.
~
c ∂t
~ ·A
For radiation, it is convenient to use the Coulomb or radiation gauge: ∇ ~ = 0, and also use
Φ = 0. Then Maxwell’s equations give us

2A(~
~ r, t) = 0, (10.97)
2
where 2 ≡ ∂ µ ∂µ = ∇2 − c12 ∂t∂
2.

Our problem is to describe A(~~ r, t) for the electromagnetic field and then to construct the electric
and magnetic fields from A(~~ r, t) using E~ = −1 ∂ A ~ and B ~ =∇~ × A.
~ Once the fields are known, the
c ∂t
field energy, field linear momentum and field angular momentum are obtained from
Z
1
Hγ = d3 r[E 2 + B 2 ]

Z
1
P~γ = d3 r[E~ × B]
~
4πc
Z
1
J~γ = ~ × B].
d3 r ~r × [E ~ (10.98)
4πc

recall that the Poynting vector S ~= 1 E ~ ×B~ gives us an energy flux density and ~g = S/c
~ 2 is the
4πc
linear moment flux density. Additional attributes of the field system, such as parity, polarization,
spin etc, will be discussed later. p
The vector potential must satisfy both the transverse condition ∇ ~ ·A
~ = 0, and the wave equation
2A(~r, t) = 0. A general decomposition of A into normal modes can now be made in which the
~ ~
above conditions are incorporated. We have, using box normalization in a volume Ω = L3 ,

~ r, t) = 11 ~ ~
X X
0 ∗
A(~ ε̂~k,α [ q~k,α (t) e+ik·~r + q~k,α (t) e−ik·~r ] . (10.100)
Ω 2
~ α=1,2
k

~ 1
The expresses the field in terms of plane waves e+ik·~r /Ω 2 with time dependent amplitudes q~k,α (t).
To assure that the vector potential, and hence the electric and magnetic fields, are real we have
o These equations involve first order derivatives in space and in time and are coupled equations. We shall see later

that this remark about the structure of the Maxwell equations probably influenced Dirac in his approach to developing
a relativistic version of Quantum Mechanics.
p The total angular momentum associated with localized, classical electromagnetic fields can be separated into

orbital plus spin parts


J~γ = J~ + J~S
ZL
1~ ~
J~L = ~ (~
d3 r A r× ∇) A
i
Z
d ~
J~S = d3 r( ~ .
A) × A (10.99)
dt
CHAPTER 10. QUANTUM DYNAMICS 284

introduced the q ∗ term, and also restricted the sum to the positive ~k values, as indicated by the
prime on the first sum. In a box of volume Ω = L3 , the allowed values of ~k are ~k = 2π L n̂, so we
consider all integers nz ≥ 0, nz = 0, 1, 2, · · · , and nx , ny = 0 ± 1, ±2, · · · , e.g., we consider the upper
hemisphere in ~n space.
The real vectors ε̂~k,α are defined for each ~k and α to assure the transverse nature of A. ~ We have
q
the following properties for these linear polarization vectors.

k̂ · ε̂~k,α = 0 for α = 1, 2
ε̂~k,α=3 = k̂ for α = 3
ε̂~k,α0 · ε̂~k,α = δα0 α (10.101)

Note for linear polarization ε̂~k,α = ε̂~∗k,α and

ε̂~k,1 × ε̂~k,2 = k̂.

Therefore for each ~k, α the corresponding term in the series, or mode, for A
~ is perpendicular to ~k
and is thus transverse in each mode, which assures that ∇ ~ ·A~ = 0.
To assure that the wave equation is satisfied the amplitude of the normal modes of the vector field
satisfy harmonic motion equations

(kc)2 q~k,α (t) + q̈~k,α (t) = 0 or q̈~k,α (t) = −ωk2 q~k,α (t), (10.102)

where ωk = kc. These are solved by q~k,α (t) = e±iωk t q~k,α , where the part q~k,α in bold face, is
independent of time. These results show it is appropriate to refer to this expansion of the vector
potential as a decomposition into normal modes. This is analogous to the expansion of modes used
in the Debye theory of specific heat.
Now the electric and magnetic fields are similarly decomposed since B ~ =∇~ ×A ~ and E
~ = −1A ~˙
c

~ r, t) 1 1 X0 X ~ ∗ ~
E(~ = − ε̂~k,α [ q̇~k,α (t) e+ik·~r + q̇~k,α (t) e−ik·~r ]
c Ω 12 α=1,2
~
k

~ r, t) i X
0
X ωk ~ ∗ ~
B(~ = 1 (k̂ × ε̂~k,α ) [ q~k,α (t) e+ik·~r − q~k,α (t) e−ik·~r ] (10.103)
Ω2 ~ α=1,2
c
k

Note both B ~ and E ~ are real.


Using these expansions of the magnetic and electric field into vector normal modes, the classical
energy for the field is given by:
1 X0 X
Hγ = {| q̇~k,α (t) |2 +ωk2 | q~k,α (t) |2 } (10.104)
4πc2 α=1,2 ~
k

Proof:
The proof involves a substitution
Z
1 ~∗ · E
~ +B~ ∗ · B]
~
Hγ = d3 r[E

1 Ω X0 X 0 1
= (ε̂~ · ε̂~ 0 0 ){2δ~k,~k0 | q̇~k,α (t) |2 }
8π Ω c2 k,α k ,α
~
k,α ~
k0 ,α0

ωk2
+ (k̂ × ε̂~k,α )(k̂ 0 × ε̂~k0 ,α0 ){2δ~k,~k0 | q~k,α (t) |2 } (10.105)
c2
and use of the following relations
Z
~ ~0
ei(k−k )·~r d3 r = Ω δ~k,~k0 (10.106)

q To describe circular polarization, one typically forms the following combinations: 1±1 (~k) ≡ ∓ √1 (ε̂~k,1 ± ε̂~k,1 )
2
and 10 (~k) = ε~k,3 .
CHAPTER 10. QUANTUM DYNAMICS 285

and Z
~ ~0
ei(k+k )·~r d3 r = Ω δ(~k + ~k 0 ) = 0, (10.107)

since both ~k and ~k 0 are restricted to the upper half plane and δ(kz + kz0 ) = 0. Also, we use

ε̂~k,α · ε̂~k,α0 = δα0 α = (k̂ × ε̂~k,α ) · (k̂ × ε̂~k,α0 ). (10.108)

We then arrive at the final form


1 X0 X
Hγ = {| q̇~k,α (t) |2 +ωk2 | q~k,α (t) |2 }. (10.109)
4πc2 α=1,2
~
k

The above form for Hγ is recognized as a sum over the energy of a set of harmonic 1-D oscillators.
p2
Recall that for the harmonic oscillator Hho = 2m + 12 mω 2 q 2 = m 2 2 2
2 [q̇ + ω q ]. Thus, the
electromagnetic field is formed by summing the energy from each mode.
Following the procedures we used for the simple harmonic oscillator, we can introduce amplitudes
that are simply related to q and q̇, but are more convenient for later generalization. the amplitudes
of the various vector normal modes of the electromagnetic field satisfy simple harmonic equations
q̈~k,α = −ωk q~k,α , and can therefore be expressed as
s
2πh̄c
q~k,α (t) = {a~ (t) + a∗−~k,α (t)} = q−

~ (t), (10.110)
| k | k,α k,α

where a~k,α (t) and a~∗k,α (t) are the positive and negative frequency solutions

a~k,α (t) = e−iωk t a~k,α


a~∗k,α (t) = e+iωk t a~∗k,α (10.111)

The reason for using a∗−~k,α (t) in the expression above for q~k,α (t) is that it produces plane waves in
opposite directions.
One has q̇~k,α (t) given by
s
2πh̄c
q̇~k,α (t) = iωk {−a~k,α (t) + a∗−~k,α (t)} = q̇−

~ (t). (10.112)
|k| k,α

These definitions simply replace the two amplitudes q±~k,α (t) by two new amplitudes a±~k,α (t) with
corresponding changes for q̇~k,α (t).
Now we have

| q̇~k,α (t) |2 = 2πh̄c2 ωk [(−a~k,α (t) + a∗−~k,α (t))(−a~∗k,α (t) + a−~k,α (t))]

and
ωk2 | q~k,α (t) |2 = 2πh̄c2 ωk [(a~k,α (t) + a∗−~k,α (t))(a ∗~k,α (t) + a−~k,α (t))].
Using these in an earlier expression for Hγ we get
1 X0
Hγ = 2h̄ωk [a~∗k,α a~k,α + a−~k,α a∗−~k,α ], (10.113)
2
~
k,α

where the time dependent phases have cancelled out. Dropping the restriction on the sum to
~k > 0, we can write the above as

1 X
Hγ = h̄ωk [a~∗k,α a~k,α + a−~k,α a∗−~k,α ],
2
~
k,α
1 X
= h̄ωk [a~∗k,α a~k,α + a~k,α a~∗k,α ], (10.114)
2
~
k,α
CHAPTER 10. QUANTUM DYNAMICS 286

since ωk = ω−k =| k | c. We leave the form above in preparation for quantizing the normal modes.
The above expression is a classical result, where the amplitudes are just numbers and so we can
write in the classical case X
Hγ = h̄ωk [a~∗k,α a~k,α ],
~
k,α

which gives the total electromagnetic energy as a sum over the energy/mode times the amount
a~∗k,α a~k,α of mode ~k, α that is excited in a particular situation.

Field Momentum
Similar steps using the classical vector normal modes apply to the field momentum P~γ . The result
is: Z
1 X
P~γ = ~ ×B
d3 r [E ~ −B~ × E]
~ = h̄~k[a~∗k,α a~k,α ]. (10.115)
8πc
~
k,α

The interpretation here is that h̄~k is the linear momentum for each particular mode and again
a~∗k,α a~k,α is the amount of that mode ~k, α excited. Giving that information is equivalent to
specifying the classical state of the free electromagnetic field system. We define the
n~k,α = a~∗k,α a~k,α =| a~k,α |2 as a set of real numbers which specify the classical state. Later we will
quantize that amount by introduction of an occupation number operator.
The treatment of the orbital and spin angular momentum of the classical and photon versions, and
other attributes will be skipped, due to time limitations.

10.6.1 Quantization of the Electromagnetic Field


~ E
So far we have simply expressed the fields (A, ~ &B)~ and the corresponding field momentum (P~ ),
field energy (Hγ ) and field angular momentum (J)~ in terms of vector normal modes. we have
found that the system can be described for the most part by giving the quantities n~k,α =| a~k,α |2 =
“amount of mode excited.” We introduce the idea of field quanta or photons by requiring that the
modes can be excited only in discrete amounts. That is, we quantize by requiring only integer
values for the numbers n~k,α r
n~k,α = 0, 1, 2, · · ·
Physically this means that we can excite a given mode only 0, 1, 2, · · · times, but not say 3/4, 4/3...
etc. times. Any k− value is allowed, any frequency and hence any color of light , but its energy
occurs in bundles of units h̄ωk , as originally stipulated by Einstein. The photon is therefore an
excitation of a vector field mode, two photons of the same ~k, α (momentum and polarization) are
represented by a double excitation of the ~k, α mode.
Quantization means there is an eigenvalue equation, expectation values and an uncertainty
principle and states represented as vectors in a Hilbert space. We have therefore the eigenvalue
equation,
N~k,α | n~k,α >= n~k,α | n~k,α >, (10.116)

for a selected mode ~k, α. Here N~k,α is an Hermitian operator in the Hilbert space | n~k,α >, which is
labelled by the real integer eigenvalue n~k,α .
We now define operators a~k,α and a~† in the Hilbert space with the properties
k,α

N~k,α = a~† a~k,α = N~†


k,α k,α

[N~k,α , a~k,α ] = −a~k,α

[ N~k,α , a~† ] = a~† , (10.117)


k,α k,α

and correspondingly,
[ a~k0 ,α0 , a~† ] = δα,α0 δ~k0 ,~k
k,α
r This is often called second quantization, but I prefer to call it the occupation number representation. After all
this is the first time we are quantizing the fields, with our “first quantization” being of the electrons.
CHAPTER 10. QUANTUM DYNAMICS 287

[ a~k0 ,α0 , a~k,α ] = 0. (10.118)


Indeed, the last two commutators, plus the definition of the number operator suffice, in that the
other relations can be deduced, as we did last term for the 1-D harmonic oscillator.
The commutator [ N~k,α , a~k,α ] = −a~k,α , tells us that a~k,α acts as a destruction (or photon
annihilation ) operator, and that a~† acts as a creation operator for photons, since we have the
k,α
properties

N~k,α (a~k,α | n~k,α >) = (n~k,α − 1)(a~k,α | n~k,α >)


N~k,α (a~† | n~k,α >) = (n~k,α + 1)(a~† | n~k,α >) (10.119)
k,α k,α

The vacuum ( or no photon n~k,α = 0 ) state is defined by N~k,α | 0 >= 0 for all ~k, α. We also have
as a direct consequence of the properties above that
q
a~† | n~k,α >= n~k,α + 1 | n~k,α + 1 >
k,α

and
a~k,α | n~k,α >= n~k,α | n~k,α − 1 > (10.120)
p

Finally, an explicit form for the | n~k,α > state is expressed in terms of the vacuum state by

1
| n~k,α > q (a~† )n~k,α | 0 > (10.121)
n~k,α ! k,α

Not only the mode ~k, α can be excited, we can have photons of other polarizations α and
frequencies. Our previous steps are now simply generalized in view of the lack of coupling between
the modes to

1
(a~†
n~k
Y
| nγ >=| {n~k1 ,α1 , n~k2 ,α2 , · · ·} >= q ) i ,αi |0>. (10.122)
ki ,αi
i=1 n~ki ,αi !

We need to state the number of photons in each mode ~ki , αi to specify a state. Here we have an
infinite product of oscillator states. Also note that the set nγ = {n~k1 ,α1 , n~k2 ,α2 , · · ·} are the
P
occupation numbers, with i n~ki ,αi giving the total number of photons of all frequencies and
polarizations.
In most cases, just a few photons are involved explicitly, so much of the notation is superfluous.
But a general state of the free electromagnetic system can be given as
X
| Ψγ >= Cnγ | nγ > (10.123)

which will be useful in the next section, when we build photon state wave packets.

10.6.2 The Uncertainty Principle For Fields


~ E,
The introduction of the operators a and a† causes the fields to be operators as well. Hence, A, ~
~
and B are now operators in Hilbert space and we have quantized these fields. Correspondingly, the
fields no longer commute and one can prove for example (as done by Jordan & Pauli) that

∂ ∂ ∂2
[Ei (~r1 t1 ), Ej (~r2 t2 )] = 4πih̄c{δij − }D(~r12 , t12 ) = [Bi (~r1 t1 ), Bj (~r2 t2 )]
c∂t1 c∂t2 ∂r1 ∂r2

[Ei (~r1 t1 ), Bj (~r2 t2 )] = 4πih̄cijk D(~r12 , t12 ) (10.124)
∂t1 ∂r2k
1
with D(~r, t) = 4πr [δ(r + ct) − δ(r − ct)]. See Gottfried pages 419-420 for prove and discussion of
these relations. The main point is that the fields are operators and that such commutators express
the uncertainty principle for fields.
CHAPTER 10. QUANTUM DYNAMICS 288

Another aspect of the quantization of fields, which we can relate to uncertainty in our ability to
measure quantities, is deduced from the commutator
r
2
~~ ] = h̄c √ 1 ε̂~ [ −a~ ei(~k·~r−ωk t) + a† e−i(~k·~r−ωk t) ].
[N~k,α , A k,α ~
2π 2 2ωk k,α k,α k,α

1 ∂ ~
= A~ . (10.125)
iωk ∂t k,α
~~ .
See the appendix below for the definition of A k,α
We can write the above commutator in terms of electric and magnetic fields as

~~ ] = − i ∂ E
[N~k,α , E ~~
k,α ωk ∂t k,α
~~ ] = − i ∂ B
[N~k,α , B ~~ . (10.126)
k,α ωk ∂t k,α
For a general state of the electromagnetic system, we have
X
| Ψγ >= C nγ | n γ > , (10.127)

where | Cnγ |2 is the probability for finding the system in state | nγ > where nγ denotes the
momentum and polarization set: {n~k1 ,α1 , n~k2 ,α2 , · · ·}. The expectation values and uncertainties in
our measurements for a general state | Ψγ >, is then given in the usual way:
q
∆N = < Ψγ | N 2 | Ψγ > − < Ψγ | N | Ψγ >2 . (10.128)

Also the uncertainty in the vector field measurement is:


q
∆Ax = < Ψγ | A2x | Ψγ > − < Ψγ | Ax | Ψγ >2 , (10.129)

with corresponding expressions for the other components and for the electric and magnetic fields.
Using the x− component case of
i ∂
[N~k,α , Ax~k,α ] = − A ~
ωk ∂t xk,α
1
and the general theorem that [A, B] = iC, implies ∆A ∆B ≥ 2 |<C>|, we get

1 1 ∂
∆N~k,α ∆Ax~k,α ≥ | <Ax~k,α>| . (10.130)
2 ωk ∂t
We can express the ratio
∆Ax~k,α
1 ∂
= ∆φ~k,α
| ωk ∂t <Ax~k,α>|
as an uncertainty in the phase and we then arrive at the phase -photon number uncertainty
principle
1
∆N~k,α ∆φ~k,α ≥ . (10.131)
2
The phase is a classical concept which is measure by comparing to some standard; if there is a
change in the phase of a signal of ∆φ due to the measurement process that change must be small
to have the classical limit. Thus the classical limit corresponds to ∆φ → 0, or
1 ∂
∆Ak  < Ak >,
ωk ∂t
and
1 ∂
∆Ek  < Ek >,
ωk ∂t
CHAPTER 10. QUANTUM DYNAMICS 289

ÈCHnLÈ2
1

0.8

0.6

0.4

0.2

nHkL
20 40 60 80 100

Figure 10.8: A wide distribution of photons.

ÈCHnLÈ2
1

0.8

0.6

0.4

0.2

nHkL
20 40 60 80

Figure 10.9: A narrow distribution of photons.

etc. That is, we must then have small fluctuations or uncertainties in the fields.
Using ∆N~k,α ∆φ~k,α ≥ 12 , in the classical limit ∆N → ∞, and the number of photons is completely
indeterminate. The wave function | ψγ > then has a wide distribution of photon numbers
For the quantum limit ∆φ → ∞, the uncertainty ∆N → 0, and the number of photons becomes
precisely known. Hence the distribution is narrow:
and | Ψγ →| n > . (In this case, < n | Ȧ | n >→ 0, and ∆φ → ∞.)
The classical limit requires not only that the phase be measurable ∆φ → 0, or ∆N~k,α  1, but
also that
∆N~k,α  <N~k,α> . (10.132)
the second condition is based on the requirement that in classical physics the energy fluctuations of
the fields be negligible compared to the average energy, i.e

∆energy = h̄ωk ∆N~k,α = fluctuations  h̄ωk <N~k,α>= average energy.

I suggest you look at page 65 in Heitler’s “Quantum Theory of Radiation,” for more discussion of
these matters.
Note that the vacuum state N~k,α | 0 >= 0 for all ~k, α has precisely zero photons. Consequently,
considerable uncertainty in the electric
P and magnetic fields occurs, which is called the vacuum
fluctuation. The zero point energy, k 12 h̄ωk arises from this residual motion. This zero point
fluctuation in the fields will play a role in spontaneous emission later.
CHAPTER 10. QUANTUM DYNAMICS 290

Appendix–The Vector Potential Field Operator


Instead of expressing A ~ in terms of the amplitude q~ , it is convenient to use the new quantity
k,α
a~k,α , which is an amplitude in the classical case, but became an operator when we quantized. The
change is as follows:

~ r, t) 1 X
0
X ~ ∗ ~
A(~ = 1 ε̂~k,α [ q~k,α (t) e+ik·~r + q~k,α (t) e−ik·~r ]
Ω 2
~ α=1,2
k
s
2πh̄c2 ~ ~
ε̂~k,α [ a~k,α ei(k·~r−ωk t) + a† ~ ei(k·~r+ωk t)
X
0
=
Ωωk − k,α
~

~ ~
+a−~k,α e−i(k·~r+ωk t) + a~† e−i(k·~r−ωk t) ]. (10.133)
k,α

With ωk = ω =| k | c, we can extend the primed sum to the full region in ~k− space and have
s
2πh̄c2 ~ ~
ε̂~k,α [ a~k,α ei(k·~r−ωk t) + a~† e−i(k·~r−ωk t) ] =
X X
~ r, t) =
A(~ ~~ (~r, t). (10.134)
ε̂~k,α A
Ωωk k,α kα
~
kα ~

The following vacuum- one photon matrix elements are particularly useful
~~ (~r, t) a†
<0|A ~~ (~r, t) | 1~ >= A
| 0 >=< 0γ | A ~~ (~r, t) (10.135)
kα ~ k,α kα k,α kα

and
~~ (~r, t) | 0 >=< 1~ | A
< 0 | a~k,α A ~~ (~r, t) | 0 >= A
~ ∗ (~r, t), (10.136)
kα k,α kα ~kα
where s
~~ (~r, t) ≡ 2πh̄c2 ~
Akα ε̂~ ei(k·~r−ωk t) .
Ωωk k,α
This is the “wave function” of a plane-wave linearly polarized photon.
For an infinite volume Ω → ∞, we use the box normalization to continuum rules
Z
X X Ω
→ d~n = d~k
n
(2π)3
~
k
s

a~k,α → a~k,α
(2π)3
~ ~0
ei(k−k )·~r (2π)3 ~ ~ 0
Z
δ~k~k0 = d~r → δ(k − k )
Ω Ω

→ δ(~k − ~k 0 ) (10.137)
(2π)3 δ~k~k0

and obtain
r
h̄c2 X d~k
Z
~ ~
~ r, t) =
A(~ √ ε̂~ [ a~k,α ei(k·~r−ωk t) + a~† e−i(k·~r−ωk t) ]. (10.138)
2π 2 α 2ωk k,α k,α

with [a~k,α , a~† 0 ] = δαα0 δ(~k − ~k 0 ).


k ,α0
CHAPTER 10. QUANTUM DYNAMICS 291

10.6.3 The Coupled System


We have discussed the free electromagnetic field, which is noninteracting. Now we introduce the
free atomic Hamiltonian plus the coupling between the atom and the electromagnetic system. As a
result of this coupling, changes occur in the atomic system, i.e. transitions occur. Also, the
coupling causes the electromagnetic system to change by alteration of the modes that are excited,
i.e, by changing the photons that are excited. The basic picture is again that of energy transfer
between two systems.
The Hamiltonian for one electron is:
~ )2
p − ec A
(~
H= + Vc (r) + ξ(r) ~` · ~s + µ ~ + H0γ ,
~s · B (10.139)
2m
where Vc is the Coulomb interaction between proton and electron. Note that the spin-dependent
terms ~` · ~s and µ ~ have to be added in by hand, i.e. by Pauli’s hands. Later, such terms will
~s · B
arise naturally from Dirac’s relativistic equation. We have
2
e ~ + e A2 + 2µ0 ~s · (∇
~ × A)
~
H = H0A + H0γ + p~ · A
mc 2mc h̄
= H0A + H0γ + V = H0 + V. (10.140)
These Hamiltonians are: the atomic Hamiltonian
p2
H0A = + Vc (r) + ξ(r) ~` · ~s,
2m
the free electromagnetic Hamiltonian:
1
h̄ωk (a~† a~kα + ),
X
H0γ =
kα 2
~

and the interaction:


2
e ~ + e A2 + 2µ0 ~s · (∇
~ × A),
~
V =+ p~ · A
mc 2mc2 h̄
using µs = gs µ0~s/h̄.
The coupling of the electron and the electromagnetic field appears in the above, since each term
has both the electron’s coordinate and momentum operator, and the creation/annihilation photon
operators.
In preparation for use of the Dirac picture and the Dyson series, note that the Schrödinger picture
field operator is the time independent operator
s
2πh̄c2 ~ ~
ε̂~ [ a~ eik·~r + a~† e−ik·~r ],
X
~ S (~r) = A(~
A ~ r, t = 0) =
Ωωk k,α k,α k,α
~

where a~† and a~k,α are time independent. To get the Dirac operator, we recall the relation
k,α
~D
between a Schrödinger and a Dirac operator ΩD (t) = eiH0 t ΩS e−iH0 t . Therefore, to generate A
we need,
a~D
k,α
(t) = eiH0 t a~k,α e−iH0 t , (10.141)
where we use a~k,α to denote the time-independent Schrödinger picture operator and a~D
k,α
(t) for the
Dirac time dependent operator, whose time dependence is
1 D
a~D ˙ (t) = [a (t), H0γ ] = −iωk a~D (t).
k,α ih̄ ~k,α k,α

Therefore
a~D
k,α
(t) = e−iωk t a~D
k,α
(0) = e−iωk t a~k,α .
So the Dirac vector field operator is
s
2πh̄c2 ~ ~
ε̂~k,α [ a~k,α ei(k·~r−ωk t) + a~† e−i(k·~r−ωk t) ],
X
~ D (~r, t) = A(~
A ~ r, t) = (10.142)
Ωωk k,α
~

CHAPTER 10. QUANTUM DYNAMICS 292

which, since it evolves in time according to H0 is called the “free field” operator. Note it is
understood that the coordinate above is also a Dirac electron location operator as
~rD (t) = eiH0A t ~r e−iH0A t and p~D (t) = eiH0A t p~ e−iH0A t . (10.143)
The states for H0 = H0A + H0γ are simply product states for the two systems, which is the general
way to express two noninteracting systems. Our earlier form H0 | n >= εn | n > now becomes
(H0A + H0γ ) | nA >| nγ >= (εnA + εnγ ) | nA >| nγ > (10.144)
for eigenstates of H0A and H0γ . For example, | nA >→| n`jm >, which are the states for the atom
and | nγ >→| 00 · · · 1~k,α , 0 · · · >=| 1~k,α > is a state for H0γ with one photon with momentum and
polarization ~k, α.
A more general description of the electromagnetic system is provided by building a “wave packet”
out of the CON set of occupation number basis states | nγ >, where nγ denotes the set of photon
occupation numbers : {n~k1 ,α1 , n~k2 ,α2 · · ·}. The wave packet is
X
| Ψγ (t) >= Cnγ (t) | nγ > . (10.145)
all nγ

These states are in general not eigenstates of H0γ and the expansion coefficients can depend on
time as shown. The wave packet can be designed to describe a classical, perhaps coherent, beam of
light which contains many photons. The above packet can describe the initial and final
electromagnetic states before and after an interaction with an atom, in which case we have photon
wave packet states | Ψiγ > and | Ψfγ (t) > and associated coefficients Cni γ and Cnf γ . Although these
photon wave packets are not orthogonal s the product states are still orthogonal if we stick to the
CON atomic states
< n0A Ψfγ | nA Ψiγ >=< n0A | nA > < Ψfγ | Ψiγ >= δn0A nA < Ψfγ | Ψiγ > . (10.146)
Our reason for considering the photon wave packets is to form a quantum mechanical description
of a classical beam of photons, which is the experimental situation for absorption and stimulated
emission processes. It will allow us to understand the connection with Einstein coefficients in black
body equilibrium.

10.6.4 Radiation
The coupling of the atom to the electromagnetic field leads to radiation, absorption, and transitions
of both systems. Our first step is to express the matrix elements of V in the Schrödinger picture
s
e X 2πh̄c2
< n0A Ψfγ | VS | nA Ψiγ > =
mc Ωωk
~

~
p · ε̂~k,α + i~s · (~k × ε̂~k,α )eik·~r ] | nA >
{< Ψfγ | a~kα | Ψiγ >< n0A | [~
~
+ < Ψfγ | a~† p · ε̂~k,α − i~s · (~k × ε̂~k,α )e−ik·~r ] | nA >
| Ψiγ >< n0A | [~

e2
+ < n0A Ψfγ | A2 (~r) | nA Ψiγ > }. (10.147)
2mc2
For more than one electron, the above result must be extended by summing over all electrons and
using atomic many-electron wave functions. The quadratic A2 term will be neglected for now; it
enters for the scattering of photons since it involves the creation and destruction (a† a terms) and
reappears in discussions of dispersion, Raman effect and diamagnetism. Having said all of that, we
will discuss the case of one-electron and will neglect the quadratic term
The time dependent perturbation theory result we obtained when n 6= n0 for a pulse starting at
t = 0, is given in lowest order by
1 t
Z
cnn0 (t) = < n | VD (τ ) | n0 > dτ. (10.148)
ih̄ 0
s One can use them to define a photon beam density matrix.
CHAPTER 10. QUANTUM DYNAMICS 293

Now we must adapt this to explain the absorption, stimulated emission, spontaneous emission and
decay and the photoelectric and ionization effects, etc. Note that | n > above denotes the product
state | nA , Ψγ > .

Absorption and Stimulated Emission- Discrete Atomic Levels


In the presence of electromagnetic radiation, an atom can absorb radiation of certain frequencies.
The atom can also be stimulated by the presence of radiation to emit radiation of certain
frequencies. (Later we shall consider the cases of spontaneous emission and the photoelectric
effect.)
figure here
The radiation consists of many photons, and in experimental set-up often consists of a pulse of
classical radiation. For example, in atomic or nuclear magnetic resonance experiments, a time
dependent B− field is felt by the atom for the time (L/v) it takes to pass through the magnetic
field, see below. In addition a perpendicular magnetic field B⊥ = B0 cos ωt is produced by coils ,
while the permanent magnet sets up the steady uniform field over a length L. A beam of radiation
can be thought of as a sequence of such pulses.
The important point is that the beam consists of many photons P which ar incoherent. A classical
beam is obtained by forming a wave packet or pulse | Ψγ >= nγ Cnγ | nγ >
Light beam figure here:
A classical beam means 1  ∆N~kα <N~kα> . Incoherent means there is no correlation between
one photon in the beam and another in the (incoherent) beam. (The basic idea is that interference
termsP between individual Pphotons inPthe classical beam can be ignored and instead of expressions
like | n ψn |2 , one has | ψn |2 → n | ψn |2 which yields a classical average over an ensemble of
photons.) These ideas will play a role in our discussion of the absorption and emission rates.
Absorption-Discrete Levels
For absorption a photon is removed and hence the a~kα part of the vector potential operator will
enter. We have:
s
e 0 f ~ D | n A Ψγ = −
i e i ωn 0 n t X 2πh̄c2
− < nA Ψγ | p~D A e A A
mc mc Ωωk
~

~
< Ψfγ | a~kα | Ψiγ > e−iωk t < n0A | p~ · ε̂~k,α eik·~r | nA >,
(10.149)

where we have dropped the spin dependent term, which can be restored later by the substitution
rule
p~ · ε̂~k,α → p~ · ε̂~k,α + i~s · (~k × ε̂~k,α ).

The a~† term does not enter because it adds a photon and also because it has a positive frequency

which would lead in the energy conservation t → ∞ limit to
R (ωn0 −ωnA +ωk )t
dτ e A → δ(ωn0A − ωnA + ωk ) = 0, by conservation of energy.
Using the above Dirac picture matrix element in our first order expression for the transition
amplitude, we have
s
e 1 X 2πh̄c2
cnA →n0A = −
mc ih̄ Ωωk
~

~
< Ψfγ | a~kα | Ψiγ >< n0A | p~ · ε̂~k,α eik·~r | nA > f (ωn0A − ωnA − ωk , t), (10.150)

which includes our familiar f (ω, t) peaked function. As in our earlier discussion, we consider the
transition rate ṗnA →n0A (t) as given by 1t | cnA →n0A |2 ,
s s
e2 1 X 2πh̄c2 2πh̄c2
ṗnA →n0A (t) = WnfAi →n0 =
A m2 c2 h̄2 Ωωk Ωωk0
~
kα;~
k0 α0
~ ~0
< n0A | p~ · ε̂~k,α eik·~r | nA > < n0A | p~ · ε̂~k0 ,α0 eik ·~r | nA >∗
CHAPTER 10. QUANTUM DYNAMICS 294

< Ψiγ | a~† 0 | Ψfγ > < Ψfγ | a~kα | Ψiγ >
k α0
f (ωn0A − ωnA − ωk , t) f ∗ (ωn0A − ωnA − ωk0 , t)
. (10.151)
t
Now we do not observe the final state of the electromagnetic field and therefore sum over all such
states. t Thus we face the “closure-” type expression
X
| Ψfγ > < Ψfγ | .
f

Here there arises a subtle point that was addressed by Glauber, who showed that the classical
states are a complete set albeit not orthonormal. Invoking that authority, we replace
X
| Ψfγ > < Ψfγ | → 1,
f

after summing over all final wave packet states. We get

· · · < Ψiγ | a~† 0 0 a~kα | Ψiγ > .


X X
WnfAi →n0 = WnA →n0A = (10.152)
A k α
f ~
kα;~
k0 α0

The expectation value < Ψiγ | a~† 0 0 a~kα | Ψiγ > ( which is a density matrix describing the initial
k α
photon beam) depends on how the beam was prepared. For a classical incoherent beam the above
average (over the ensemble of photons) satisfies

< Ψiγ | a~† 0 0 a~kα | Ψiγ >= δα0 α δ~k0~k < Ψiγ | a~† a~kα | Ψiγ >= δα0 α δ~k0~k n̄~kα . (10.153)
k α kα

Here n̄~kα denotes the average number of photons of momentum ~k and polarization α in the
classical incoherent beam. We shall also assume that the photons, within a solid angle dΩk̂ , are all
in the same direction; thus
n̄~kα = n̄α (ωk )δ(k̂ − k̂0 ) (10.154)

for a beam in the direction k̂0 . Also


Z
dΩk̂ n̄~kα = n̄α (ωk ).
∆Ω

Using these properties of the beam, the “transition rate” or probability rate for absorption is
ṗnA →n0A (t) as given by 1t | cnA →n0A |2 ,

e2 1 X 2πh̄c2 ~
WnA →n0A = |< n0A | p~ · ε̂~k,α eik·~r | nA >|2
m2 c2 h̄2 Ωωk
~

| f (ωn0A − ωnA − ωk , t) |2
n̄α (ωk )δ(k̂ − k̂0 ) . (10.155)
t
Since, using box normalization conventions:
Z ∞
X L 3X ~ Ω
∆~n = ( ) ∆k = k 2 dkdΩk̂
2π (2π)3 0
~
n ~
k
R∞ R∞ dωk
and 0
dk → 0 c , we have
∞ | f (ωn0A − ωnA − ωk , t) |2 ∞
| f (ω, t) |2
Z Z
dωk = dωk = 2π. (10.156)
0 t −∞ t
t The initial state could be included in the sum over final states since it can not be reached by the destruction

operator any way.


CHAPTER 10. QUANTUM DYNAMICS 295

~
We can also factor out the n̄α (ωk )k 2 |< n0A | p~ · ε̂~k,α eik·~r | nA >|2 terms since they are smooth
broad functions of k. We arrive at:
e2 X 1 ωk ~
WnA →n0A = 2 2 |< n0A | p~ · ε̂~k,α eik·~r | nA >|2 n̄α (ωk ), (10.157)
m c α 2π h̄c
with h̄ωk = h̄(ωn0A − ωnA ).
For a given polarization α of the incident beam the transition rate is
e2 h̄ωk n̄α (ωk ) ~ · ε̂~ ei~k·~r | nA >|2
WnαA →n0 = 2 3
|< n0A | ∇ k,α (10.158)
A m c 2π
Intensity of the Beam
The intensity of the photon beam is the quantity, along with the direction and the polarization,
that characterizes the beam. To obtain an expression for the intensity, we consider the average
energy u
h̄ωk < Ψiγ | a~† a~kα | Ψiγ >
X
< Ψiγ | H | Ψiγ > =

~

X
= h̄ωk n̄α (ωk )
~

Ω X ∞
Z
= dkk 2 h̄ωk n̄α (ωk ). (10.159)
(2π)3 α 0

The energy flux [ energy/(area-sec) ] is c× the energy per volume; recall j = ρc. Therefore, we
deduce from the above that the energy flux [ energy/(area-sec) ] is
< Ψiγ | H | Ψiγ > 1 X ∞ ωk2 X Z ∞
Z
c = dω k 2 h̄ω n̄
k α (ωk ) = dωk Iα (ωk ), (10.160)
Ω (2π)3 α 0 c α 0

where we have introduced the intensity of the beam for a given polarization Iα (ωk ), which has the
units of energy/area
1 ωk2
Iα (ωk ) ≡ h̄ωk n̄α (ωk ). (10.161)
(2π)3 c2
The combination Iα (ωk )dωk is the flux of energy in the interval dωk and has the units of energy /(
area-sec).
Using this expression for the beam intensity, the result for the absorption of a photon by a system
( say an atom or a nucleus) is for a beam of polarization α
e2 Iα (ωk ) ~ · ε̂~ ei~k·~r | nA >|2
WnαA →n0 = (2π)2 |< n0A | ∇
A m2 c ωk2 k,α

~
2 Iα (ωk ) 1
p · ε̂~k,α + i~s · (~k × ε̂~k,α )]eik·~r | nA > 2
< n0A | [~
= α0 (2π) | | (10.162)
h̄ωk ωk m
2
e
Here α0 = h̄c = 1/137, is the fine structure constant and we have reintroduced the spin terms. The
units for this result are W = (energy/area)(1/energy) sec (area/sec)2 = 1/sec= probabilty/sec.

Stimulated and Spontaneous Emission


A similar procedure can be applied to the case of an incident state | nA Ψiγ > leading to a lower
final state | n0A Ψfγ > with the addition or emission of a photon.
The basic change is that the creation operator part of A ~ now enters the transition amplitude with
† −iωk t
a corresponding change to a a and a e to obtain:
s
e 1 X 2πh̄c2
cnA →n0A = −
mc ih̄ Ωωk
~

~
< Ψfγ | a~† | Ψiγ >< n0A | p~ · ε̂~k,α e−ik·~r | nA > f (ωn0A − ωnA + ωk , t),(10.163)

1
u Note that the term
P
2
h̄ωk = E∞ , is an infinite amount of zero point energy which should be on both sides and
therefore cancels–we use an energy scale “relative” to this troublesome zero point energy.
CHAPTER 10. QUANTUM DYNAMICS 296

with the infinite time limit giving ωk = ωnA − ωn0A . Most of the steps are the same as before with
one very important exception, the order of the operators changes from the absorption case of a† a
to the order aa† for the emission case and we see

< Ψiγ | a~kα a~† 0 | Ψfγ >= δαα0 δ~k~k0 (n̄~kα + 1),
k α0

where the n̄ part leads to stimulated emission and the “one” part gives spontaneous emission. The
emission, which occurs for no incident radiation (n̄~kα = 0) and is called spontaneous emission,
arises from the fluctuations in the electric and magnetic fields in the vacuum (no photon) state.
Even when there are no photons present, there are ground state vibrations of 12 h̄ω~kα for each mode
of the electromagnetic field. Clearly, the “one” term above arises from the non-commutativity of
aa† = a† a + 1, which is essentially the uncertainty principle for our quantized fields.
For an incident incoherent classical beam of polarization α the resulting emission rate is
e2 h̄ωk
Z
WnαA →n0 (emission) = 2 3 ~ · ε̂~ e−i~k·~r | nA >|2 ,
dΩk̂ (n̄~kα + 1) |< n0A | ∇ (10.164)
A m c 2π kα

which consists of two terms

WnαA →n0 (emission) = WnαA →n0 (stimulated) + WnαA →n0 (spontaneous).


A A A

The first term depends on the presence of incident radiation and is called stimulated emission. For
this term we can complete the angle integration as done before with a beam direction included and
using the same steps as before with
Z
dΩk̂ n̄~kα = n̄α (ωk ).

The result for stimulated radiation is therefore:


e2 h̄ωk n̄α (ωk ) ~ · ε̂~ e−i~k·~r | nA >|2
WnαA →n0 (stimulated) = |< n0A | ∇ k,α
A m2 c3 2π
2
e Iα (ωk ) ~ · ε̂~ e−i~k·~r | nA >|2
= (2π)2 |< n0A | ∇
m2 c ωk2 k,α

~
2 Iα (ωk ) 1
p · ε̂~k,α − i~s · (~k × ε̂~k,α )]e−ik·~r | nA > 2
< n0A | [~
= α0 (2π) | | ,
h̄ωk ωk m
(10.165)

where the last step includes the spin term and expresses the transition rate in terms of the incident
e2
beam intensity. Recall α0 = h̄c , is the fine structure constant the units for this result are as for the
prior absorption case W = (energy/area)(1/energy) sec (area/sec)2 = 1/sec= probabilty/sec. Note
that this expression is very similar to that for absorption of a photon, which will prove to be
essential for an understanding of black-body radiation. Indeed, it can be shown, from the above
integral, that the rate for stimulated emission and for absorption between the same two level ( and
the same α) are related by

WnαA →n0 (stimulated emission) = Wnα0 →nA (absorption). (10.166)


A A

This is seen here by an integration by parts (see HW). This relationship is called detailed balance
and is also a consequence of “time reversal invariance” or “invariance under the reversal of motion.”
Spontaneous Emission
The spontaneous emission rate is given by
e2 h̄ωk ~ · ε̂~ e−i~k·~r | nA >|2 dΩ ,
dWnαA →n0 = |< n0A | ∇ k,α k̂ (10.167)
A m2 c3 2π
where the initial photon state | Ψiγ > equal to the photon vacuum can be used. This transition
probability can also be written to include the spin term as
dWnαA →n0 h̄ 2 ωk 1 ~
A
= α0 ( ) p · ε̂~k,α − i~s · (~k × ε̂~k,α )]e−ik·~r | nA >|2 .
|< n0A | [~ (10.168)
dΩk̂ mc 2π h̄2
CHAPTER 10. QUANTUM DYNAMICS 297

e2
Note that we again have the fine structure constant α0 = h̄c , and now the Compton wave length of
dWnα 0
A →nA

the electron mc =λ–c appears. The units of the above are 1/(sec-steradian). Note that
dΩk̂
gives the probability per sec for emission into a solid angle dΩk̂ ,
For spontaneous emission we can rewrite the above as
dWnαA →n0
A –2c ωk 1 |< n0A | [∇
= α0 λ ~ · ε̂~ − i ~s · (~k × ε̂~ )]e−i~k·~r | nA >|2 . (10.169)
dΩk̂ 2π k,α h̄ k,α

In these expressions, the integral over the solid angle dΩk̂ can not in general be done, as was the
case with a directed incident beam with n̄~k,α = nα (ωk )δ(k̂ − k̂0 ). Here we express the total
probability/sec for spontaneous emission simply as
Z dWnαA →n0
WnαA →n0 = dΩk̂ A
(10.170)
A dΩk̂

Spontaneous Emission -Multipoles and Selection Rules


The probability/sec for an atom to spontaneously emit a photon of momentum ~k and polarization
α into a solid angle dΩk̂ is give by
dWnαA →n0
A
= α0 λ ~ · ε̂~ − i ~s · (~k × ε̂~ )]e−i~k·~r | nA >|2 .
–2c ωk 1 |< n0A | [∇ (10.171)
dΩk̂ 2π k,α h̄ k,α

Multipole Expansion
~
Now let us examine the e−ik·~r factor which we shall expand. The matrix element involves an
integration over atomic wave functions < n0A | ~r > and < ~r | nA > which are significant only inside
the size of the atom for both n0A and nA states. Therefore, for transitions between discrete levels
(bounded), the value of ~r inside the integral is r ≤ atomic size ≈ A0 /Z = Bohr radius divided by
Z of the atom.
The value of k of the photon is deduced from conservation of energy.
e2 2 1 1
h̄ωk = h̄ck = εnA − εn0A = − Z ( 2 − 0 2 ) > 0. (10.172)
2A0 nA nA
2 2 2
e e Z α0 Z Z
Therefore h̄ck ≤ 2A 0
Z 2 . Hence kr ≤ h̄c 2A0 r ≈ 2 ≈ 300 , for atoms. For transitions in light
atoms ( Z small and optical transitions) kr ≈ 1/300, which is quite small. For higher Z values and
say X-ray transitions, the value of kr increases and one might begin to need higher terms in the
~
expansion of e−ik·~r . For π µ and antiprotonic atoms, one might also get higher order terms. For
nuclei, higher order terms do occur.
For atomic systems at usual densities, it suffices to use the estimate
kr ≈ Z/300
as small and to approximate
~
e−ik·~r ≈ 1.
Higher order terms in this expansion −i(~k · ~r + · · · lead to higher multipoles such as magnetic
dipole (M1), and electric quadrupole(E2) transitions. v
If we postpone consideration of the spin term, the probability per second for spontaneous emission
of a photon of energy h̄ωk and polarization α into the solid angle dΩk̂ is given in the “Electric
Dipole Approximation” by
dWnαA →n0 1
A –2c ωk
= α0 λ |< n0A | p~ · ε̂~k,α 1 | nA >|2 . (10.173)
dΩk̂ 2πh̄2
~ = e~r by using
An additional simplification is to introduce the dipole operator D
i mi p2
p~ = m [H0A , ~r] = [ , ~r], (10.174)
h̄ h̄ 2m
v These correction terms also include the so-called retardation corrections to lower multipoles.
CHAPTER 10. QUANTUM DYNAMICS 298

2
p
which is clearly valid for H0A = 2m + Vc (r). If H0A includes a spin orbit interaction and/or any
nonlocal terms further discussion is needed. It turns out that the following discussion is largely
still valid ( see Siegert’s theorem).
For our present case then, the above commutator tells us that

< n0A | p~ | nA >→ imωk < n0A | ~r | nA > . (10.175)

With that substitution, we arrive at the result


dWnαA →n0 e2 ωk3
A
= |< n0A | ~r · ε̂~k,α | nA >|2 . (10.176)
dΩk̂ h̄c3 2π

which makes contact with Heisenberg’s use of electric dipole radiation, combined with the
Rayleigh-Ritz combination principle to discover matrix mechanics in the first place. So the cycle
has been completed in part.
The cycle could be completed by returning to Einstein’s approach to black-body radiation. The
point is that he considered the equality of the absorption and stimulated emission and also
included the spontaneous emission. The equilibrium of an atom in a radiation field then led him to
Plank’s law, where the course now ends.

10.7 The Dirac Equation


10.7.1 References on the Dirac Equation
The following are good references for learning about the Dirac equation:
• P.A.M. Dirac, The Prediction of Antimatter, the first H.R. Crane lecture at University of
Michigan, Ann Arbor (1978).
• ”Origins of Dirac’s Electrons 1925-1028”, by D.F/Moyer, American Journal of Physics, 49 10
(1981) History outline
• ”Reminiscences about a great physicist- Paul Adrien Maurice Dirac,” edited by B.N.
Kursunoglu and E’P Wigner, Cambridge Press (1987).
• H. Feshbach and F. Villars, Rev. Mod. Phys, 30, 24 (1958)-list of earlier references plus a
good rendition of the development of basic ideas.

• H. A. Bethe and R. Jackiw, ”Intermediate Quantum Mechanics.” A consise, insightful


rendition in Bethe’s wonderful style
• J. D. Bjorken and S.D. Drell, ”Relativistic Quantum Mechanics,” (I) and ”Relativistic
Fields,” (II).

• A Messiah, ” Quantum Mechanics,” Chapter 20 vol. II.


• E Merzbacher, ” Quantum Mechanics,” Chapter 23.
• S. Schweber, ”An Introduction to relativistic Quantum Field Theory,” very readable.

• H. A. Bethe and E.E. Salpeter , ”Quantum mechanics of One- and Two- Electron Atoms,” a
classic.
• W. Pauli, ”Lectures on Physics,” vol 6. A classic.
• P.A.M. Dirac, ”The Priciples of Quantum mechanics,” of course!

• Itzykson and Zuber ”Quantum Field Theory.” Has a lot of good material.
CHAPTER 10. QUANTUM DYNAMICS 299

10.7.2 The Dirac Equation


This is a subject of great beauty and much significance. If you have not previously studied the
Dirac equation, i am pleased to be the person to expose you to it. If you have already studied this
subject, then you will appreciate that one can look at the Dirac equation many times and continue
to learn from it. Indeed, Dirac himself has expressed wonder at the content of his own creation; he
had not expected to discover the subtle relationship between spin and relativity, nor did he expect
to discover the existence of antimatter ( see his attached lecture).
References that I recommend to you are listed above, including several papers that present the
historical background. Unfortunately, we have little time for this great subject, so I’ll try to be
efficient.
The main goal is to combine quantum with relativity theory. and to obtain a logically consistent
theory of the relativistic, quantum motion of particles. Recall that in quantum mechanics we
described the motion of a particle by a wave function Ψ(~r, t), from which we obtain the positive
probability Ψ∗ Ψ of finding the particle at ~r at time t. As the particle moves, the function Ψ(~r, t)
which is an expression of the uncertainty principle. For example, we have a wave packet for a free
particle
~
ei(k·~r−ωk t) ~ ~
Z
Ψ(~r, t) = φ(k − k0 ) d3 k, (10.177)
(2π)3/2
and we take
~
h̄2 k 2 ei(k·~r−ωk t) ~ ~
Z

(ih̄ − H0 )Ψ(~r, t) = (h̄ωk − ) φ(k − k0 ) d3 k = 0, (10.178)
∂t 2m (2π)3/2
p2 h̄2 ~ For a free particle h̄ωk = h̄2 k 2
where for non-relativistic motion H0 = 2m = − 2m ∇2 , and p~ = h̄i ∇. 2m
k . Here we incorporate the Einstein E = h̄ω and DeBroglie p~ = h̄~k relations. Thus,
h̄ 2
or ωk = 2m
h̄ 2
with ωk = 2m k the above wave packet describes a moving and spreading probability distribution
w
and the dynamical equation is simply

H0 Ψ(~r, t) = h̄i Ψ(~r, t), (10.179)
∂t
Schrödinger’s equation, which is a complex diffusion equation involving first time and second
spatial derivatives. To include potentials we take

h̄2 2
H = H0 + V = − ∇ + V. (10.180)
2m
How can we generalize the above discussion to include relativity? One possible attack is to consider
p p
E = h̄ωk = p2 c2 + m2 c4 = h̄2 k 2 c2 + m2 c4 , (10.181)

E = h̄ω and p~ = h̄~k for each Fourier component. That suggests taking
using p
H0 = −h̄2 c2 ∇2 + m2 c4 and

∂ p
− −h̄2 c2 ∇2 + m2 c4 }Ψ(~r, t) =
{ih̄
∂t
~
ei(k·~r−ωk t) ~ ~
Z p
(h̄ωk − p2 c2 + m2 c4 ) φ(k − k0 ) d3 k = 0, (10.182)
(2π)3/2
p
with ωk = p2 c2 + m2 c4 . hence we again have Schrödinger’s equation

H0 Ψ(~r, t) = h̄i Ψ(~r, t), (10.183)
∂t
p
but with H0 = −h̄2 c2 ∇2 + m2 c4 . The wave function Ψ(~r, t) is again a moving and spreading
wave packet.
w Recall the wave packet discussion last term and its relationship with the uncertainty principle
CHAPTER 10. QUANTUM DYNAMICS 300

Have we solved the problem of combining relativity and quantum theory, including the
transformation theory that relates momentum and space descriptions? Not really, since: (a) the
above operator is awkward and although defined readily in momentum space, it is nonlocal,ugly
and difficult to generalize to include electromagnetism in coordinate space; and (b) the time and
space aspects are not on an equal footing as expected for a relativistic theory; and (c) it is difficult
to obtain an equation of continuity for a positive definite probability using the above H0 . Recall
h̄2 ∂
that for − 2m ∇2 Ψ(~r, t) = h̄i ∂t Ψ(~r, t), we obtained

∂ ~ · ~j
ρ = −∇ (10.184)
∂t
with ρ = Ψ∗ Ψ > 0 and ~j = 2imh̄ ~ − Ψ∇Ψ
(Ψ∗ ∇Ψ ~ ∗ ), which allows us to interpret Ψ∗ Ψ as a
p
h̄2
probability density and ~j as a probability current for H0 = − 2m ∇2 . For H0 = −h̄2 c2 ∇2 + m2 c4 .
the above steps do not apply and thus this is not a valid solution to the problem of combining
relativity and quantum mechanics.
let us try starting with E 2 = p2 c2 + m2 c4 or
∂ 2
[{ih̄ } − {−h̄2 c2 ∇2 + m2 c4 }]Ψ(~r, t) =
∂t
~
ei(k·~r−ωk t) ~ ~
Z
[h̄2 ωk2 − (p2 c2 + m2 c4 )] φ(k − k0 ) d3 k = 0. (10.185)
(2π)3/2
This seems valid and we now have
∂2
[−h̄2 c2 ∇2 + h̄2 + m2 c4 ]Ψ(~r, t) = 0. (10.186)
∂t2
This equation was original proposed by DeBroglie and then extended to include the
electromagnetic field by Schrödinger, who did not publish it because it gave the wrong spectrum
for Hydrogen(see Dirac’s 1978 Crane lecture). It was later published by Klein and Gordon and is
called the Klein-Gordon (KG) equation, which including the vector and scalar electromagnetic
fields is
∂ h̄ ~ e~ 2
[(ih̄ − eΦ)2 − ( ∇ − A) h̄] Ψ(~r, t) = m2 c4 Ψ(~r, t). (10.187)
∂t i c
Have we obtained the right solution now? Not quite; although the KG equation looked promising
for a while. It certainly treats time and space on an equal footing, but it has a serious problem.
Consider again the continuity equation derivation:

∂2
h̄2 Ψ(~r, t) = h̄2 c2 ∇2 Ψ(~r, t) − m2 c4 Ψ(~r, t)
∂t2
∂2
h̄2 2 Ψ∗ (~r, t) = h̄2 c2 ∇2 Ψ∗ (~r, t) − m2 c4 Ψ∗ (~r, t)
∂t
2
∗ ∂ ∂2 ~ · (Ψ∗ ∇Ψ
~ − Ψ∇Ψ
~ ∗ ).
2
h̄ (Ψ 2
Ψ − Ψ 2 Ψ∗ = h̄2 c2 ∇ (10.188)
∂t ∂t

We obtain a continuity equation ∂t ~ · ~j with the same definition for the probability current
ρ = −∇
~j = h̄ ∗~ ~ ∗
2im (Ψ ∇Ψ − Ψ∇Ψ ), but with

ih̄ ∂ ∂
ρ= (Ψ∗ Ψ − Ψ Ψ∗ ). (10.189)
2mc2 ∂t ∂t
For a complete system we have
Z Z Z
∂ 3 3 ~ ~ ~ · ~j → 0,
ρd r = − d r∇ · j = − dS (10.190)
∂t

on an infinite surface. Therefore, the value of d3 rρ is a constant for a complete system, which is
R

nice, but what about interpreting ρ as a positive definite probability? This ρ is certainly
normalizable and real, but it is not positive definite.
CHAPTER 10. QUANTUM DYNAMICS 301

~
r −ωt)
i(k·~
To see this problem consider one solution of the KG p equation Ψ = e /(2π)3/2 for which
E 2 = (h̄ω)2 = h̄2 c2 k 2 + m2 c4 . The energy E = ± h̄2 c2 k 2 + m2 c4 has plus and minus solutions.
ih̄ ∗ E ∗
For this case we have ρ = 2mc 2 (−iω)2Ψ Ψ = mc2 Ψ Ψ, from which we see that ρ > 0 for E > 0,

and ρ < 0 for E < 0. The fact that ρ can be negative led some to interpret ρ as a charge density,
rather than as a positive probability density, but that is not a sufficient solution.
∂2 ∂2
The basic problem is that the KG equation involves ∂t 2 as well as ∂x2 and hence requires

specifying both Ψ(~r, t) as well as ∂t Ψ(~r, t) at t = 0 to obtain the subsequent wave function. x We
would like to have a theory which requires that specifying the wave function at an initial time
dictates its future y and that has a positive definite probability density. It must also treat space
and time on an equal footing. At this stage Dirac enters (1928)!

Some additional comments on the Klein-Gordon equation


The KG equation was shown by Pauli and Weiskopf (1934) to be a valid equation for a scalar field,
i.e. it is not an equation for a state vector but for a scalar field. It later turns out that the
one-particle dirac equation develops difficulties that lead to an interpretation of it also as an
equation for a spin 1/2 field rather than as an equation for a state vector. In these subsequent
developments (largely due to Dirac), the KG is then an equation for spin zero fields, whereas the

Dirac equation is an equation for spin 1/2 fields, and the state vector has H | Ψ >= ih̄ ∂t | Ψ >,
dynamics.
It is also possible to turn the second order KG equation into coupled first order equations and then
deal with two boundary conditions on a two component state vector:

φ
Ψ≡( ),
χ

where φ and χ are specified at t = 0. Thus one can avoid the issue of having to specify both and at
t = 0, which is a helpful step that entails introducing multicomponent wave functions. Pauli took
such a step, but Dirac when even deeper.

10.8 Derivation of the Dirac Equation


We want to avoid negative probability densities and still treat space and time on an equal footing.
This consideration, plus the wish to preserve the transformation theory, led Dirac to the overall
form
∂ ∂ ∂ ∂
∝ [ + βmc − αx + αy + αz ]Ψ(~r, t) = 0 (10.191)
∂t ∂x ∂y ∂z
To recover the relationship E 2 = p2 c2 + m2 c4 , Dirac assumed that the quantities β, αx , αy , αz were
N × N matrices, for example as
 x
α11 · · · α1N
αx →  ... ..  (10.192)

··· . 
αN 1 ··· αN N

and that Ψ was not a single function but a column matrix of length N × 1
 
Ψ1
 Ψ2 
Ψ→ .  (10.193)
 
.
 . 
ΨN

~ ≡ îαx + ĵαy + k̂αz ) the above can be written as


Using the notation α
1 ∂ ~ nm Ψm + i mc βnm Ψm = 0,
~ ·∇
Ψn + α (10.194)
c ∂t h̄
x It is the second derivative in time in the KG that yields the first time derivative in the expression for ρ.
y as ∂
in the original Schrödinger equation, which has it i ∂t term.
CHAPTER 10. QUANTUM DYNAMICS 302

where a sum over the repeated index m is understood and the factor of 1c and the factor mc h̄ were
introduced for consistent units (α and β are unitless). The role of the “i ” will be seen later. This
equation can be written with matrix indices suppressed as
1 ∂ ~ + i mc βΨ = 0,
~ · ∇Ψ
Ψ+α (10.195)
c ∂t h̄
or in great detail as
N N N N
1 ∂ X
x ∂ X
y ∂ X
z ∂ mc X
Ψn + αnm Ψm + αnm Ψm + αnm Ψm + i βnm Ψm = 0. (10.196)
c ∂t m=1
∂x m=1
∂y m=1
∂z h̄ m=1

Clearly, the abbreviated form is preferable.


So far all we know is that α~ , β are four matrices of size N × N and that Ψ is a N × 1 column
matrix.
Note that we ultimately want to get a wave equation to describe free particle motion. The fact
that in the Dirac equation we deal with a first order time derivative, rather than a second order
time derivative is similar to the Maxwell equations situation. Maxwell’s equations involve only first
order time and spatial derivatives, but the fields have components and obey coupled equations in
~ ×E
such a way as to yield a wave equation for those fields, i.e. ∇ ~ = − 1 ∂ B~ and ∇
~ ×B~ = 1 ∂ E~
c ∂t c ∂t
2
combine to give the wave equations 2E ~ = 0 and 2B ~ = 0 where 2 ≡ ∇2 − 12 ∂ 2 .
c ∂t
Thus there is hope that Dirac’s coupled equations will give a description of particle motion; indeed,
the extra degrees of freedom on Ψm will yield the spin and the existence of antiparticles.
We now seek a positive probability density ρ >, . and a properly defined current ~j. Start from
1 ∂ ~ + i mc βΨ = 0,
~ · ∇Ψ
Ψ+α (10.197)
c ∂t h̄
we can take the complex conjugate
1 ∂ ∗ ∗ ~ ∗m − i mc βnm

Ψ + (~
αnm · ∇)Ψ Ψ∗m = 0, (10.198)
c ∂t n h̄
and form the 1 × N adjoint matrix

Ψ† = Ψ∗1 Ψ∗N

··· . (10.199)

Now introduce the adjoints of the “Dirac matrices ”


∗ T∗ †
α
~ nm = α
~ mn =α
~ mn
∗ T∗ †
βnm = βmn = βmn (10.200)

and we can write


1 ∂ ∗ ~ ∗m ) · α † mc ∗ †
Ψn + (∇Ψ ~ mn −i Ψ β = 0. (10.201)
c ∂t h̄ m mn
Combining this adjoint equation with the original form Dirac deduced that
1 ∂ 1 ∂ 1 ∂
(Ψ∗n Ψn ) = ( Ψ∗n )Ψn + Ψ∗n ( Ψn ) (10.202)
c ∂t c ∂t c ∂t

View publication stats

You might also like