Quantum Mechanics I - Temple Math [PDF]

Aug 29, 2013 - 3.7.24 The Continuity Equation and Particle Conservation . . . 169. 4 Applications of Quantum Mechanics.

0 downloads 13 Views 4MB Size

Recommend Stories


Quantum Mechanics I
Open your mouth only if what you are going to say is more beautiful than the silience. BUDDHA

PDF Quantum Mechanics
We must be willing to let go of the life we have planned, so as to have the life that is waiting for

PHY 212: Quantum Mechanics I
When you do things from your soul, you feel a river moving in you, a joy. Rumi

Quantum mechanics
You miss 100% of the shots you don’t take. Wayne Gretzky

[PDF] Introduction to Quantum Mechanics
So many books, so little time. Frank Zappa

PDF Principles of Quantum Mechanics
What you seek is seeking you. Rumi

quantum mechanics
Nothing in nature is unbeautiful. Alfred, Lord Tennyson

Quantum Mechanics
Stop acting so small. You are the universe in ecstatic motion. Rumi

quantum mechanics
Live as if you were to die tomorrow. Learn as if you were to live forever. Mahatma Gandhi

Quantum mechanics
Kindness, like a boomerang, always returns. Unknown

Idea Transcript


Quantum Mechanics I Peter S. Riseborough August 29, 2013

Contents 1 Principles of Classical Mechanics 1.1 Lagrangian Mechanics . . . . . . . . . . . . . . . . . . . 1.1.1 Exercise 1 . . . . . . . . . . . . . . . . . . . . . . 1.1.2 Solution 1 . . . . . . . . . . . . . . . . . . . . . . 1.1.3 The Principle of Least Action . . . . . . . . . . . 1.1.4 The Euler-Lagrange Equations . . . . . . . . . . 1.1.5 Generalized Momentum . . . . . . . . . . . . . . 1.1.6 Exercise 2 . . . . . . . . . . . . . . . . . . . . . . 1.1.7 Solution 2 . . . . . . . . . . . . . . . . . . . . . . 1.2 Hamiltonian Mechanics . . . . . . . . . . . . . . . . . . 1.2.1 The Hamilton Equations of Motion . . . . . . . . 1.2.2 Exercise 3 . . . . . . . . . . . . . . . . . . . . . . 1.2.3 Solution 3 . . . . . . . . . . . . . . . . . . . . . . 1.2.4 Time Evolution of a Physical Quantity . . . . . . 1.2.5 Poisson Brackets . . . . . . . . . . . . . . . . . . 1.3 A Charged Particle in an Electromagnetic Field . . . . . 1.3.1 The Electromagnetic Field . . . . . . . . . . . . 1.3.2 The Lagrangian for a Classical Charged Particle 1.3.3 Exercise 4 . . . . . . . . . . . . . . . . . . . . . . 1.3.4 Solution 4 . . . . . . . . . . . . . . . . . . . . . . 1.3.5 The Hamiltonian of a Classical Charged Particle 1.3.6 Exercise 5 . . . . . . . . . . . . . . . . . . . . . . 1.3.7 Solution 5 . . . . . . . . . . . . . . . . . . . . . .

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

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

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

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

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

9 9 10 10 12 15 16 17 17 19 20 21 21 22 22 25 25 26 27 27 28 29 29

2 Failure of Classical Mechanics 2.1 Semi-Classical Quantization . 2.1.1 Exercise 6 . . . . . . . 2.1.2 Solution 6 . . . . . . . 2.1.3 Exercise 7 . . . . . . . 2.1.4 Solution 7 . . . . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

33 34 34 34 36 36

. . . . .

1

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

3 Principles of Quantum Mechanics 3.1 The Principle of Linear Superposition . . . . . . . 3.2 Wave Packets . . . . . . . . . . . . . . . . . . . . . 3.2.1 Exercise 8 . . . . . . . . . . . . . . . . . . . 3.2.2 Solution 8 . . . . . . . . . . . . . . . . . . . 3.2.3 Exercise 9 . . . . . . . . . . . . . . . . . . . 3.2.4 Solution 9 . . . . . . . . . . . . . . . . . . . 3.3 Probability, Mean and Deviations . . . . . . . . . . 3.3.1 Exercise 10 . . . . . . . . . . . . . . . . . . 3.3.2 Solution 10 . . . . . . . . . . . . . . . . . . 3.3.3 Exercise 11 . . . . . . . . . . . . . . . . . . 3.3.4 Solution 11 . . . . . . . . . . . . . . . . . . 3.3.5 Exercise 12 . . . . . . . . . . . . . . . . . . 3.3.6 Solution 12 . . . . . . . . . . . . . . . . . . 3.4 Operators and Measurements . . . . . . . . . . . . 3.4.1 Operator Equations . . . . . . . . . . . . . 3.4.2 Operator Addition . . . . . . . . . . . . . . 3.4.3 Operator Multiplication . . . . . . . . . . . 3.4.4 Commutators . . . . . . . . . . . . . . . . . 3.4.5 Exercise 13 . . . . . . . . . . . . . . . . . . 3.4.6 Solution 13 . . . . . . . . . . . . . . . . . . 3.4.7 Exercise 14 . . . . . . . . . . . . . . . . . . 3.4.8 Solution 14 . . . . . . . . . . . . . . . . . . 3.4.9 Exercise 15 . . . . . . . . . . . . . . . . . . 3.4.10 Solution 15 . . . . . . . . . . . . . . . . . . 3.4.11 Exercise 16 . . . . . . . . . . . . . . . . . . 3.4.12 Solution 16 . . . . . . . . . . . . . . . . . . 3.4.13 Exercise 17 . . . . . . . . . . . . . . . . . . 3.4.14 Solution 17 . . . . . . . . . . . . . . . . . . 3.4.15 Exercise 18 . . . . . . . . . . . . . . . . . . 3.4.16 Solution 18 . . . . . . . . . . . . . . . . . . 3.4.17 Eigenvalue Equations . . . . . . . . . . . . 3.4.18 Exercise 19 . . . . . . . . . . . . . . . . . . 3.4.19 Exercise 20 . . . . . . . . . . . . . . . . . . 3.4.20 Solution 20 . . . . . . . . . . . . . . . . . . 3.4.21 Exercise 21 . . . . . . . . . . . . . . . . . . 3.4.22 Solution 21 . . . . . . . . . . . . . . . . . . 3.4.23 Exercise 22 . . . . . . . . . . . . . . . . . . 3.4.24 Solution 22 . . . . . . . . . . . . . . . . . . 3.4.25 Adjoint or Hermitean Conjugate Operators 3.4.26 Hermitean Operators . . . . . . . . . . . . . 3.4.27 Exercise 23 . . . . . . . . . . . . . . . . . . 3.4.28 Exercise 24 . . . . . . . . . . . . . . . . . . 3.4.29 Solution 24 . . . . . . . . . . . . . . . . . . 3.4.30 Exercise 25 . . . . . . . . . . . . . . . . . . 3.4.31 Solution 25 . . . . . . . . . . . . . . . . . . 2

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

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

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

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

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

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

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

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

38 40 43 46 46 48 49 52 54 54 57 57 58 59 60 61 62 62 64 66 66 67 67 68 68 70 70 71 71 72 72 73 75 75 76 78 78 79 79 81 85 86 86 86 89 89

3.5

3.4.32 Exercise 26 . . . . . . . . . . . . . . . . . . . . . . . . . 3.4.33 Solution 26 . . . . . . . . . . . . . . . . . . . . . . . . . 3.4.34 Eigenvalues and Eigenfunctions of Hermitean Operators 3.4.35 Exercise 27 . . . . . . . . . . . . . . . . . . . . . . . . . 3.4.36 Solution 27 . . . . . . . . . . . . . . . . . . . . . . . . . 3.4.37 Exercise 28 . . . . . . . . . . . . . . . . . . . . . . . . . 3.4.38 Solution 28 . . . . . . . . . . . . . . . . . . . . . . . . . 3.4.39 Exercise 29 . . . . . . . . . . . . . . . . . . . . . . . . . 3.4.40 Solution 29 . . . . . . . . . . . . . . . . . . . . . . . . . 3.4.41 Exercise 30 . . . . . . . . . . . . . . . . . . . . . . . . . 3.4.42 Solution 30 . . . . . . . . . . . . . . . . . . . . . . . . . 3.4.43 Exercise 31 . . . . . . . . . . . . . . . . . . . . . . . . . 3.4.44 Exercise 32 . . . . . . . . . . . . . . . . . . . . . . . . . 3.4.45 Solution 32 . . . . . . . . . . . . . . . . . . . . . . . . . 3.4.46 Hermitean Operators and Physical Measurements . . . . 3.4.47 Exercise 33 . . . . . . . . . . . . . . . . . . . . . . . . . 3.4.48 Solution 33 . . . . . . . . . . . . . . . . . . . . . . . . . 3.4.49 Exercise 34 . . . . . . . . . . . . . . . . . . . . . . . . . 3.4.50 Solution 34 . . . . . . . . . . . . . . . . . . . . . . . . . 3.4.51 Exercise 35 . . . . . . . . . . . . . . . . . . . . . . . . . 3.4.52 Solution 35 . . . . . . . . . . . . . . . . . . . . . . . . . Quantization . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.5.1 Relations between Physical Operators . . . . . . . . . . 3.5.2 The Correspondence Principle . . . . . . . . . . . . . . 3.5.3 The Complementarity Principle . . . . . . . . . . . . . . 3.5.4 Coordinate Representation . . . . . . . . . . . . . . . . 3.5.5 Momentum Representation . . . . . . . . . . . . . . . . 3.5.6 Exercise 36 . . . . . . . . . . . . . . . . . . . . . . . . . 3.5.7 Exercise 37 . . . . . . . . . . . . . . . . . . . . . . . . . 3.5.8 Exercise 38 . . . . . . . . . . . . . . . . . . . . . . . . . 3.5.9 Solution 38 . . . . . . . . . . . . . . . . . . . . . . . . . 3.5.10 Exercise 39 . . . . . . . . . . . . . . . . . . . . . . . . . 3.5.11 Solution 39 . . . . . . . . . . . . . . . . . . . . . . . . . 3.5.12 Exercise 40 . . . . . . . . . . . . . . . . . . . . . . . . . 3.5.13 Solution 40 . . . . . . . . . . . . . . . . . . . . . . . . . 3.5.14 Exercise 41 . . . . . . . . . . . . . . . . . . . . . . . . . 3.5.15 Solution 41 . . . . . . . . . . . . . . . . . . . . . . . . . 3.5.16 Commuting Operators and Compatibility . . . . . . . . 3.5.17 Non-Commuting Operators . . . . . . . . . . . . . . . . 3.5.18 Exercise 42 . . . . . . . . . . . . . . . . . . . . . . . . . 3.5.19 Solution 42 . . . . . . . . . . . . . . . . . . . . . . . . . 3.5.20 The Uncertainty Principle . . . . . . . . . . . . . . . . . 3.5.21 Exercise 43 . . . . . . . . . . . . . . . . . . . . . . . . . 3.5.22 Solution 43 . . . . . . . . . . . . . . . . . . . . . . . . . 3.5.23 Exercise 44 . . . . . . . . . . . . . . . . . . . . . . . . . 3.5.24 Solution 44 . . . . . . . . . . . . . . . . . . . . . . . . . 3

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

91 91 92 94 94 95 96 99 99 99 100 102 102 103 104 105 105 107 107 108 109 112 112 112 113 114 117 123 125 125 126 127 127 128 128 129 129 130 132 132 132 133 135 135 136 136

3.6

3.7

3.5.25 Exercise 45 . . . . . . . . . . . . . . . . . . . . . . . 3.5.26 Solution 45 . . . . . . . . . . . . . . . . . . . . . . . 3.5.27 Exercise 46 . . . . . . . . . . . . . . . . . . . . . . . 3.5.28 Solution 46 . . . . . . . . . . . . . . . . . . . . . . . The Philosophy of Measurement . . . . . . . . . . . . . . . 3.6.1 Exercise 47 . . . . . . . . . . . . . . . . . . . . . . . 3.6.2 Solution 47 . . . . . . . . . . . . . . . . . . . . . . . 3.6.3 Exercise 48 . . . . . . . . . . . . . . . . . . . . . . . 3.6.4 Solution 48 . . . . . . . . . . . . . . . . . . . . . . . 3.6.5 Exercise 49 . . . . . . . . . . . . . . . . . . . . . . . 3.6.6 Solution 49 . . . . . . . . . . . . . . . . . . . . . . . Time Evolution . . . . . . . . . . . . . . . . . . . . . . . . . 3.7.1 The Schr¨ odinger Picture. . . . . . . . . . . . . . . . 3.7.2 Exercise 50 . . . . . . . . . . . . . . . . . . . . . . . 3.7.3 Solution 50 . . . . . . . . . . . . . . . . . . . . . . . 3.7.4 The Heisenberg Picture. . . . . . . . . . . . . . . . . 3.7.5 Exercise 51 . . . . . . . . . . . . . . . . . . . . . . . 3.7.6 Solution 51 . . . . . . . . . . . . . . . . . . . . . . . 3.7.7 Exercise 52 . . . . . . . . . . . . . . . . . . . . . . . 3.7.8 Solution 52 . . . . . . . . . . . . . . . . . . . . . . . 3.7.9 Exercise 53 . . . . . . . . . . . . . . . . . . . . . . . 3.7.10 Exercise 54 . . . . . . . . . . . . . . . . . . . . . . . 3.7.11 Solution 54 . . . . . . . . . . . . . . . . . . . . . . . 3.7.12 Exercise 55 . . . . . . . . . . . . . . . . . . . . . . . 3.7.13 Exercise 56 . . . . . . . . . . . . . . . . . . . . . . . 3.7.14 The Schr¨ odinger Equation . . . . . . . . . . . . . . . 3.7.15 Exercise 57 . . . . . . . . . . . . . . . . . . . . . . . 3.7.16 Solution 57 . . . . . . . . . . . . . . . . . . . . . . . 3.7.17 Time Development of a Wave Packet . . . . . . . . . 3.7.18 Exercise 58 . . . . . . . . . . . . . . . . . . . . . . . 3.7.19 Solution 58 . . . . . . . . . . . . . . . . . . . . . . . 3.7.20 Time Evolution and Energy Eigenfunctions . . . . . 3.7.21 Exercise 59 . . . . . . . . . . . . . . . . . . . . . . . 3.7.22 Solution 59 . . . . . . . . . . . . . . . . . . . . . . . 3.7.23 The Correspondence Principle . . . . . . . . . . . . 3.7.24 The Continuity Equation and Particle Conservation

4 Applications of Quantum Mechanics 4.1 Exact Solutions in One Dimension . . . . . . . . . . . . . 4.1.1 Particle Confined in a Deep Potential Well . . . . 4.1.2 Time Dependence of a Particle in a Deep Potential 4.1.3 Exercise 60 . . . . . . . . . . . . . . . . . . . . . . 4.1.4 Particle Bound in a Shallow Potential Well . . . . 4.1.5 Exercise 61 . . . . . . . . . . . . . . . . . . . . . . 4.1.6 Solution 61 . . . . . . . . . . . . . . . . . . . . . . 4.1.7 Scattering from a Shallow Potential Well . . . . . 4

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

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

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

136 137 137 137 140 143 144 145 146 147 148 149 150 151 151 152 153 153 154 154 155 155 155 156 157 157 160 160 161 162 162 164 166 166 168 169

. . . . . . Well . . . . . . . . . . . . . . .

. . . . . . . .

174 174 174 182 183 183 190 191 194

4.2

4.3 4.4

4.1.8 Exercise 62 . . . . . . . . . . . . . . . . . . . . 4.1.9 Solution 62 . . . . . . . . . . . . . . . . . . . . 4.1.10 Exercise 63 . . . . . . . . . . . . . . . . . . . . 4.1.11 Solution 63 . . . . . . . . . . . . . . . . . . . . 4.1.12 The Threshold Energy for a Bound State . . . 4.1.13 Transmission through a Potential Barrier . . . 4.1.14 Exercise 64 . . . . . . . . . . . . . . . . . . . . 4.1.15 Solution 64 . . . . . . . . . . . . . . . . . . . . 4.1.16 The Double Well Potential . . . . . . . . . . . 4.1.17 The delta function Potential . . . . . . . . . . . 4.1.18 Bound States of a delta function Potential . . . 4.1.19 Exercise 65 . . . . . . . . . . . . . . . . . . . . 4.1.20 Solution 65 . . . . . . . . . . . . . . . . . . . . 4.1.21 Exercise 66 . . . . . . . . . . . . . . . . . . . . 4.1.22 Solution 66 . . . . . . . . . . . . . . . . . . . . 4.1.23 Exercise 67 . . . . . . . . . . . . . . . . . . . . 4.1.24 Solution 67 . . . . . . . . . . . . . . . . . . . . 4.1.25 Exercise 68 . . . . . . . . . . . . . . . . . . . . 4.1.26 Solution 68 . . . . . . . . . . . . . . . . . . . . The One-Dimensional Harmonic Oscillator . . . . . . . 4.2.1 The Raising and Lowering Operators . . . . . . 4.2.2 The Effect of the Lowering Operator . . . . . . 4.2.3 The Ground State . . . . . . . . . . . . . . . . 4.2.4 The Effect of The Raising Operator . . . . . . 4.2.5 The Normalization . . . . . . . . . . . . . . . . 4.2.6 The Excited States . . . . . . . . . . . . . . . . 4.2.7 Exercise 69 . . . . . . . . . . . . . . . . . . . . 4.2.8 Solution 69 . . . . . . . . . . . . . . . . . . . . 4.2.9 Exercise 70 . . . . . . . . . . . . . . . . . . . . 4.2.10 Solution 70 . . . . . . . . . . . . . . . . . . . . 4.2.11 Exercise 71 . . . . . . . . . . . . . . . . . . . . 4.2.12 Time Development of the Harmonic Oscillator 4.2.13 Exercise 72 . . . . . . . . . . . . . . . . . . . . 4.2.14 Solution 72 . . . . . . . . . . . . . . . . . . . . 4.2.15 Hermite Polynomials . . . . . . . . . . . . . . . 4.2.16 Exercise 73 . . . . . . . . . . . . . . . . . . . . 4.2.17 Solution 73 . . . . . . . . . . . . . . . . . . . . 4.2.18 Exercise 74 . . . . . . . . . . . . . . . . . . . . 4.2.19 Solution 74 . . . . . . . . . . . . . . . . . . . . 4.2.20 The Completeness Condition . . . . . . . . . . Dual-symmetry . . . . . . . . . . . . . . . . . . . . . . Bargmann Potentials . . . . . . . . . . . . . . . . . . . 4.4.1 Exercise 75 . . . . . . . . . . . . . . . . . . . . 4.4.2 Solution 75 . . . . . . . . . . . . . . . . . . . . 4.4.3 Exercise 76 . . . . . . . . . . . . . . . . . . . . 4.4.4 Solution 76 . . . . . . . . . . . . . . . . . . . . 5

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

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

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

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

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

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

199 199 201 201 203 204 207 208 209 213 215 224 224 226 226 232 232 234 235 237 238 238 239 240 241 241 243 244 246 246 246 247 249 250 251 256 256 257 257 258 260 263 266 266 270 270

4.5

4.6

4.4.5 Exercise 77 . . . . . . . . . . . . . . . . . . 4.4.6 Solution 77 . . . . . . . . . . . . . . . . . . Orbital Angular Momentum . . . . . . . . . . . . . 4.5.1 Exercise 78 . . . . . . . . . . . . . . . . . . 4.5.2 Solution 78 . . . . . . . . . . . . . . . . . . 4.5.3 Exercise 79 . . . . . . . . . . . . . . . . . . 4.5.4 Simultaneous Eigenfunctions. . . . . . . . . 4.5.5 The Raising and Lowering Operators . . . . 4.5.6 The Eigenvalues and Degeneracy . . . . . . 4.5.7 The Effect of the Raising Operators. . . . . 4.5.8 Explicit Expressions for the Eigenfunctions 4.5.9 Legendre Polynomials . . . . . . . . . . . . 4.5.10 Associated Legendre Functions . . . . . . . 4.5.11 Spherical Harmonics . . . . . . . . . . . . . 4.5.12 Exercise 80 . . . . . . . . . . . . . . . . . . 4.5.13 Solution 80 . . . . . . . . . . . . . . . . . . 4.5.14 Exercise 81 . . . . . . . . . . . . . . . . . . 4.5.15 Solution 81 . . . . . . . . . . . . . . . . . . 4.5.16 Exercise 82 . . . . . . . . . . . . . . . . . . 4.5.17 Solution 82 . . . . . . . . . . . . . . . . . . 4.5.18 The Addition Theorem . . . . . . . . . . . 4.5.19 Finite-Dimensional Representations . . . . 4.5.20 Exercise 83 . . . . . . . . . . . . . . . . . . 4.5.21 Exercise 84 . . . . . . . . . . . . . . . . . . 4.5.22 Solution 84 . . . . . . . . . . . . . . . . . . 4.5.23 The Laplacian Operator . . . . . . . . . . . 4.5.24 An Excursion into d-Dimensional Space . . 4.5.25 Exercise 85 . . . . . . . . . . . . . . . . . . 4.5.26 Solution 85 . . . . . . . . . . . . . . . . . . Spherically Symmetric Potentials . . . . . . . . . . 4.6.1 Exercise 86 . . . . . . . . . . . . . . . . . . 4.6.2 Solution 86 . . . . . . . . . . . . . . . . . . 4.6.3 The Free Particle . . . . . . . . . . . . . . . 4.6.4 The Spherical Square Well . . . . . . . . . . 4.6.5 Exercise 87 . . . . . . . . . . . . . . . . . . 4.6.6 Solution 87 . . . . . . . . . . . . . . . . . . 4.6.7 Exercise 88 . . . . . . . . . . . . . . . . . . 4.6.8 Exercise 89 . . . . . . . . . . . . . . . . . . 4.6.9 Solution 89 . . . . . . . . . . . . . . . . . . 4.6.10 Exercise 90 . . . . . . . . . . . . . . . . . . 4.6.11 Solution 90 . . . . . . . . . . . . . . . . . . 4.6.12 Exercise 91 . . . . . . . . . . . . . . . . . . 4.6.13 Solution 91 . . . . . . . . . . . . . . . . . . 4.6.14 Exercise 92 . . . . . . . . . . . . . . . . . . 4.6.15 Solution 92 . . . . . . . . . . . . . . . . . . 4.6.16 Ladder operators for a free particle . . . . . 6

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

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

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

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

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

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

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

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

273 273 276 278 278 280 282 284 285 286 287 290 293 294 298 298 302 303 304 305 306 308 312 312 313 315 317 319 320 323 323 324 325 334 342 343 345 345 345 346 346 348 348 349 349 351

4.7

4.8

4.6.17 The Rayleigh Equation . . . . . . . . . . . 4.6.18 The Isotropic Planar Harmonic Oscillator . 4.6.19 The Spherical Harmonic Oscillator . . . . . 4.6.20 Exercise 93 . . . . . . . . . . . . . . . . . . 4.6.21 Solution 93 . . . . . . . . . . . . . . . . . . 4.6.22 Exercise 94 . . . . . . . . . . . . . . . . . . 4.6.23 The Bound States of the Coulomb Potential 4.6.24 Exercise 95 . . . . . . . . . . . . . . . . . . 4.6.25 Exercise 96 . . . . . . . . . . . . . . . . . . 4.6.26 Solution 96 . . . . . . . . . . . . . . . . . . 4.6.27 Exercise 97 . . . . . . . . . . . . . . . . . . 4.6.28 Ladder Operators for the Hydrogen Atom . 4.6.29 Rydberg Wave Packets . . . . . . . . . . . . 4.6.30 Laguerre Polynomials . . . . . . . . . . . . 4.6.31 Exercise 98 . . . . . . . . . . . . . . . . . . 4.6.32 Solution 98 . . . . . . . . . . . . . . . . . . 4.6.33 Exercise 99 . . . . . . . . . . . . . . . . . . 4.6.34 Solution 99 . . . . . . . . . . . . . . . . . . 4.6.35 Exercise 100 . . . . . . . . . . . . . . . . . . 4.6.36 Solution 100 . . . . . . . . . . . . . . . . . A Charged Particle in a Magnetic Field . . . . . . 4.7.1 Exercise 101 . . . . . . . . . . . . . . . . . . 4.7.2 Exercise 102 . . . . . . . . . . . . . . . . . . 4.7.3 Solution 102 . . . . . . . . . . . . . . . . . 4.7.4 The Degeneracy of the Landau Levels . . . 4.7.5 Exercise 103 . . . . . . . . . . . . . . . . . . 4.7.6 Solution 103 . . . . . . . . . . . . . . . . . 4.7.7 The Aharonov-Bohm Effect . . . . . . . . . The Pauli Spin Matrices . . . . . . . . . . . . . . . 4.8.1 Exercise 104 . . . . . . . . . . . . . . . . . . 4.8.2 Solution 104 . . . . . . . . . . . . . . . . . 4.8.3 Exercise 105 . . . . . . . . . . . . . . . . . . 4.8.4 Solution 105 . . . . . . . . . . . . . . . . . 4.8.5 Exercise 106 . . . . . . . . . . . . . . . . . . 4.8.6 Solution 106 . . . . . . . . . . . . . . . . . 4.8.7 Exercise 107 . . . . . . . . . . . . . . . . . . 4.8.8 Solution 107 . . . . . . . . . . . . . . . . . 4.8.9 Exercise 108 . . . . . . . . . . . . . . . . . . 4.8.10 Solution 108 . . . . . . . . . . . . . . . . . 4.8.11 The Pauli Equation . . . . . . . . . . . . . 4.8.12 Spin Dynamics . . . . . . . . . . . . . . . . 4.8.13 Exercise 109 . . . . . . . . . . . . . . . . . . 4.8.14 Solution 109 . . . . . . . . . . . . . . . . . 4.8.15 Exercise 110 . . . . . . . . . . . . . . . . . . 4.8.16 Solution 110 . . . . . . . . . . . . . . . . . 4.8.17 The Berry Phase . . . . . . . . . . . . . . . 7

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

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

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

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

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

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

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

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

355 358 360 362 363 365 367 374 374 374 375 376 380 383 390 390 392 393 395 395 399 400 400 400 401 403 404 406 412 415 415 417 417 419 419 420 421 423 424 426 428 430 430 431 432 434

4.9

Transformations and Invariance . . . . . 4.9.1 Time Translational Invariance . . 4.9.2 Translational Invariance . . . . . 4.9.3 Periodic Translational Invariance 4.9.4 Exercise 111 . . . . . . . . . . . . 4.9.5 Solution 111 . . . . . . . . . . . 4.9.6 Rotational Invariance . . . . . . 4.9.7 Exercise 112 . . . . . . . . . . . . 4.9.8 Solution 112 . . . . . . . . . . . 4.9.9 Exercise 113 . . . . . . . . . . . . 4.9.10 Solution 113 . . . . . . . . . . . 4.9.11 Exercise 114 . . . . . . . . . . . . 4.9.12 Solution 114 . . . . . . . . . . . 4.9.13 Gauge Invariance . . . . . . . . . 4.9.14 Exercise 116 . . . . . . . . . . . . 4.9.15 Solution 116 . . . . . . . . . . . 4.9.16 Galilean Boosts . . . . . . . . . .

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

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

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

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

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

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

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

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

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

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

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

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

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

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

5 The Rotating Planar Oscillator 6 Dirac Formulation 6.1 Dirac Notation . . . . . . . . . . . . . . . 6.1.1 Bracket Notation . . . . . . . . . . 6.1.2 Operators . . . . . . . . . . . . . . 6.1.3 Adjoints and Hermitean Operators 6.1.4 Representation of Operators . . . . 6.2 Representations . . . . . . . . . . . . . . . 6.3 Gram-Schmidt Orthogonalization . . . . . 7 Appendices

438 439 440 443 450 451 452 461 461 463 463 468 468 471 472 473 474 476

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

482 482 483 484 485 486 487 489 490

8

1

Principles of Classical Mechanics

Newton’s laws can be reformulated in a variety of different ways. These different formulations provide more powerful and elegant methods for solving problems which involve many different variables and have a natural formulation in non-Cartesian coordinate systems, such as spherical polar coordinates. In nonCartesian coordinate systems, vector formulations are complicated by the fact that the orthogonal directions associated with the variables depend upon the values of the generalized coordinates. The advantage of the alternate formulation of Newton’s laws is based on the fact that they involve scalar quantities rather than vector quantities, therefore, they do not require transforming the equations of motion between the Cartesian coordinates and the non-Cartesian system.

1.1

Lagrangian Mechanics

The Lagrangian approach to classical mechanics is based on a scalar quantity, the Lagrangian L, which depends upon the generalized coordinates and velocities. For a Cartesian coordinate system, the coordinates are the position of the particle x, y and z. The generalized velocities are the time derivatives of the coordinates, which we represent by x, ˙ y˙ and z. ˙ For a non-Cartesian coordinate system, such as spherical polar coordinates, the generalized coordinates for one particle are r, θ, and ϕ. The generalized ˙ and ϕ. velocities are the time derivatives of the coordinates, which are r, ˙ θ, ˙ The Lagrangian is given by the difference of the kinetic energy T and the potential energy V , L = T − V (1)

In Cartesian coordinates, we have L =

m 2 m 2 m 2 x˙ + y˙ + z˙ − V (x, y, z) 2 2 2

(2)

In spherical polar coordinates, we have L =

m r˙ 2 m r2 θ˙2 m r2 sin2 θ ϕ˙ 2 + + − V (r, θ, ϕ) 2 2 2

9

(3)

z

r

θ r y ϕ x

Figure 1: The Spherical Polar Coordinate System. A general point is labelled by the coordinates (r, θ, ϕ). ——————————————————————————————————

1.1.1

Exercise 1

Find the Lagrangian for a particle in terms of spherical polar coordinates. ——————————————————————————————————

1.1.2

Solution 1

The Lagrangian for a particle in a potential is given by m 2 r˙ − V (r) 2

(4)

∂r dr ∂r dθ ∂r dϕ + + ∂r dt ∂θ dt ∂ϕ dt

(5)

L = and with r ≡ (r, θ, ϕ) one has r˙ =

but the three orthogonal unit vectors of spherical polar coordinates eˆr , eˆθ and eˆϕ are defined as the directions of increasing r, increasing θ and increasing ϕ. 10

z

dr er r dθ θ eθ θ

dθ θ

r y dϕ ϕ ϕ

r sinθ θ dϕ ϕ eϕ

x

Figure 2: The Spherical Polar Coordinate System. An orthogonal set of unit vectors eˆr , eˆθ and eˆϕ can be constructed which, respectively, correspond to the directions of increasing r, θ and ϕ. Thus, ∂r ∂r ∂r ∂θ ∂r ∂ϕ

= eˆr = r eˆθ = r sin θ eˆϕ (6)

Hence

dr dθ dϕ + r eˆθ + r sin θ eˆϕ dt dt dt and as the unit vectors are orthogonal   2  2  2  m dr dθ dϕ L = + r2 + r2 sin2 θ − V (r) 2 dt dt dt r˙ = eˆr

(7)

(8)

——————————————————————————————————

11

For a problem involving N particles, we denote the generalized coordinates by qi , where i runs over the 3N values corresponding to the 3 coordinates for each of the N particles, and the generalized velocities by q˙i . The Lagrangian L is a function of the set of qi and q˙i , and we shall write this as L(qi , q˙i ) in which only one set of coordinates and velocities appears. However, L depends on all the coordinates and velocities. The Lagrangian is the sum of the kinetic energy of the particles minus the total potential energy, which is the sum of the external potentials acting on each of the particles together with the sum of any interaction potentials acting between pairs of particles.

1.1.3

The Principle of Least Action

The equations of motion originate from an extremum principle, often called the principle of least action. The central quantity in this principle is given by the action S which is a number that depends upon the specific function which is a trajectory qi (t0 ). These trajectories run from the initial position at t0 = 0, which is denoted by qi (0), to a final position at t0 = t, denoted by qi (t). These two sets of values are assumed to be known, and they replace the two sets of initial conditions, qi (0) and q˙i (0), used in the solution of Newton’s laws. There are infinitely many arbitrary trajectories that run between the initial and final positions. The action for any one of these trajectories, qi (t0 ) is given by a number which has the value of the integration Z t S = dt0 L(qi (t0 ), q˙i (t0 )) (9) 0

The value of S depends on the particular choice of trajectory qi (t0 ). The action is an example of a functional S[qi (t0 )] as it yields a number that depends upon the choice of a function. The extremum principle asserts that the value of S is an extremum, i.e. a maximum, minimum or saddle point, for the trajectory which satisfies Newton’s laws. To elucidate the meaning of the extremal principle, we shall consider an arbitrary trajectory qi (t0 ) that goes between the initial and final position in a time interval of duration t. Since this trajectory is arbitrary, it is different from the trajectory that satisfies Newton’s laws, which as we shall show later is an extremal trajectory qiex (t0 ). The difference or deviation between the arbitrary trajectory and the extremal trajectory is defined by δqi (t0 ) = qi (t0 ) − qiex (t0 )

(10)

An important fact is that this deviation tends to zero at the end points t0 = 0 and t0 = t since our trajectories are defined to all run through the specific initial qi (0) and final positions qi (t) at t0 = 0 and t0 = t. Let us consider the variety of the plots of δqi (t0 ) versus t0 . There are infinitely many different curves. Let us concentrate on a single shape of the curve, then we can generate 12

6

q(tf) 5

q(t)

q(t)

4

ex

q (t)

3

2

q(ti)

1

0 -1

0

ti

1

t

2

3

tf

Figure 3: Arbitrary trajectories qi (t) originating from a specific initial point qi (ti ) at t = ti and ending up at a specific final point qi (tf ) at time t = tf . a whole family of such curves by either increasing or decreasing the magnitude of the deviation by a factor of λ. The family of trajectories is given by qi (t0 ) = qiex (t0 ) + λ δqi (t0 )

(11)

When λ = 0 the original curve reduces to the extremal curve and when λ = 1 we recover our initial choice for the arbitrary trajectory. If we substitute this family of trajectories into the action, we would find a number S(λ) that depends on λ. This function S(λ) should be extremal, i.e. either a maximum, minimum, or a saddle point as a function of λ at λ = 0 if the action is an extremum at the extremal trajectory. The condition that the action is extremal is just that ∂S = 0 (12) ∂λ at λ = 0, or the first order term in the Taylor series expansion of S(λ) in λ is zero. Let us first look at a simple example of motion in one dimension, where the Lagrangian is given by L(x, x) ˙ =

m 2 x˙ − V (x) 2 13

(13)

6

q(tf)

δq(t)

5

q(t)

q(t)

4

ex

q (t)

3

2

q(ti)

1

0 -1

0

ti

1

t

2

3

tf

Figure 4: Arbitrary trajectories qi (t) going between specific initial and final points, and the extremal trajectory qiex (t). The deviation δqi (t) is defined as δqi (t) = qi (t) − qiex (t). and let us substitute x(t0 ) = xex (t0 ) + λ δx(t0 )

(14)

in S(λ) and expand in powers of λ, Z S(λ)

t

dt0 L(xex (t0 ) + λ δx(t0 ) , x˙ ex (t0 ) + λ δ x(t ˙ 0 ))

= 0

Z

t

dt0 L(xex (t0 ), x˙ ex (t0 ))   Z t ∂ ∂ 0 0 0 0 0 0 + λ dt L(x(t ), x(t ˙ )) δx(t ) + L(x(t ), x(t ˙ )) δ x(t ˙ 0) ∂x ∂ x˙ 0 λ=0 λ=0 =

0

+ O(λ2 )

(15)

In the above expression, the partial derivatives of the Lagrangian are evaluated with the extremal trajectory. Since we are only concerned with the condition that S is extremal at λ = 0, the higher order terms in the Taylor expansion in λ are irrelevant. If S is to be extremal, then the extremum condition means that the term linear in λ must vanish, no matter what our particular choice of

14

δx(t0 ) is. Thus, we require that   Z t ∂ ∂ 0 0 0 0 0 0 0 L(x(t ), x(t ˙ )) δx(t ) + L(x(t ), x(t ˙ )) δ x(t ˙ ) = 0 dt ∂x ∂ x˙ 0 λ=0 λ=0 (16) for any shape of δx(t0 ). Since this expression involves both δx(t0 ) and δ x(t ˙ 0 ), we shall eliminate the time derivative of the deviation in the second term. To do this we integrate the second term by parts, that is Z t 0 0 0 ∂ dt L(x(t ), x(t ˙ )) δ x(t ˙ 0) = ∂ x ˙ 0 λ=0 t   Z t d ∂ ∂ 0 0 L(x(t0 ), x(t ˙ 0 )) δx(t0 ) − dt0 L(x(t ), x(t ˙ )) δx(t0 ) ∂ x˙ dt0 ∂ x˙ λ=0

0

0

λ=0

(17) The boundary terms vanish at the beginning and the end of the time interval since the deviations δx(t0 ) vanish at both these times. On substituting the expression (17) back into the term of S(λ) linear in λ, one obtains    Z t d ∂ ∂ 0 0 0 0 0 dt L(x(t ), x(t ˙ )) − L(x(t ), x(t ˙ )) δx(t0 ) = 0 0 ∂x ∂x dt ˙ 0 λ=0 λ=0 (18) This integral must be zero for all shapes of the deviation δx(t0 ) if xex (t0 ) is the extremal trajectory. This can be assured if the term in the square brackets is identically zero. This gives the equation   d ∂ ∂ 0 0 0 0 L(x(t ), x(t ˙ )) − L(x(t ), x(t ˙ )) = 0 (19) ∂x dt0 ∂ x˙ λ=0 λ=0 which determines the extremal trajectory. Now using the form of the Lagrangian given in equation 13, one finds ∂V (xex (t0 )) d x(t ˙ 0) + m = 0 ∂x dt0

(20)

which is identical to the equations found from Newton’s laws. Thus, the extremal principle reproduces the results obtained from Newton’s laws.

1.1.4

The Euler-Lagrange Equations

Let us now go back to the more general case with N particles, and arbitrary coordinates qi and arbitrary Lagrangian L. It is straight forward to repeat the derivation of the extremal condition and find that the equations of motion for the extremal trajectory qi (t0 ) reduce to the 3N equations,   ∂ d ∂ 0 0 0 0 L(qj (t ), q˙j (t )) − L(qj (t ), q˙j (t )) = 0 (21) ∂qi dt0 ∂ q˙i 15

6

x(tf) 5

δx(t)

x(t)

4

3

xex(t)

2

x(ti) 1

ti

tf

0 -1

0

1

2

3

t

Figure 5: Arbitrary trajectories x(t) going between specific initial and final points, and the extremal trajectory xex (t). The deviation δx(t) is defined as δx(t) = x(t) − xex (t). where there is one equation for each value of i. The value of j is just the dummy variable which reminds us that L depends on all the coordinates and velocities. These equations are the Euler-Lagrange equations, and are a set of second order differential equations which determine the classical trajectory.

1.1.5

Generalized Momentum

The angular momentum is an example of what we call a generalized momentum. We define a generalized momentum in the same way as the components of momentum are defined for a particle in Cartesian coordinates. The generalized momentum pi conjugate to the generalized coordinate qi is given by the equation   ∂L pi = (22) ∂ q˙i Thus, in a Cartesian coordinate system, we find the x-component of a particle’s momentum is given by   ∂L px = ∂ x˙ = mx˙ (23) 16

Likewise, for the y and z components 

py

∂L ∂ y˙ = my˙







=

(24)

and pz

∂L ∂ z˙ = mz˙ =

(25)

The Euler-Lagrange equations of motion for the general case is re-written in terms of the generalized momentum as   d pi ∂L − = 0 (26) ∂qi dt0 This equation has the same form as Newton’s laws involving the rate of change of momentum on one side and the derivative of the Lagrangian w.r.t a coordinate on the other side. ——————————————————————————————————

1.1.6

Exercise 2

Find the classical equations of motion for a particle, in spherical polar coordinates. ——————————————————————————————————

1.1.7

Solution 2

The generalized momenta are found via pr

=



=



=

∂L = m r˙ ∂ r˙ ∂L = m r2 θ˙ ∂ θ˙ ∂L = m r2 sin2 θ ϕ˙ ∂ ϕ˙

17

(27)

The Euler-Lagrange equations become   dpr ∂L ∂V = = m r θ˙2 + sin2 θ ϕ˙ 2 − dt ∂r ∂r dpθ ∂L ∂V = = m r2 sin θ cos θ ϕ˙ 2 − dt ∂θ ∂θ dpϕ ∂L ∂V = = − dt ∂ϕ ∂ϕ

(28)

——————————————————————————————————

18

1.2

Hamiltonian Mechanics

Hamiltonian Mechanics formulates mechanics not in terms of the generalized coordinates and velocities, but in terms of the generalized coordinates and momenta. The Hamiltonian will turn out to be the equivalent of energy. Since Newton’s laws give rise to a second order differential equation and require two initial conditions, to solve Newton’s laws we need to integrate twice. The first integration can be done with the aid of an integrating factor. For example, with mx ¨ = −

∂V ∂x

(29)

the integrating factor is the velocity, x. ˙ On multiplying the equation by the integrating factor and then integrating, one obtains m 2 x˙ = E − V (x) 2

(30)

where the constant of integration is the energy E. Note that the solution is now found to lie on the surface of constant energy in the two-dimensional space formed by x and x. ˙ The solution of the mechanical problem is found by integrating once again. The point is, once we have obtained the energy, we are closer to finding a solution of the equations of motion. Hamiltonian mechanics results in a set of first order differential equations. The Hamiltonian, H(qi , pi , t), is a function of the generalized coordinates qi and generalized momentum pi . It is defined as a Legendre transformation of the Lagrangian X H(qi , pi , t) = q˙i pi − L(qi , q˙i , t) (31) i

The Legendre transformation has the effect of eliminating the velocity q˙i and replacing it with the momentum pi . The equations of motion can be determined from the Lagrangian equations of motion. Since the Hamiltonian is considered to be a function of coordinates and momenta alone, an infinitesimal change in H occurs either through an infinitesimal change in the coordinates dqi , momenta dpi or, if the Lagrangian has any explicit time dependence, through dt,  X  ∂H ∂H ∂H dqi + dpi + dt (32) dH = ∂q ∂p ∂t i i i However, from the definition of H one also has  X  ∂L ∂L ∂L dH = q˙i dpi + pi dq˙i − dq˙i − dqi − dt ∂ q˙i ∂qi ∂t i

19

(33)

The terms proportional to dq˙i cancel as, by definition, pi is the same as ∂∂L q˙i . Thus, we have  X  ∂L ∂L dH = q˙i dpi − dqi − dt (34) ∂qi ∂t i The cancellation of the terms proportional to the infinitesimal change dq˙i is a result of the Legendre transformation, and confirms that the Hamiltonian is a function of only the coordinates and momenta. We also can use the Lagrangian ∂L equations of motion to express ∂q as the time derivative of the momentum p˙i . i

1.2.1

The Hamilton Equations of Motion

We can now compare the specific form of the infinitesimal change in H found above, with the infinitesimal differential found from its dependence on pi and qi . On equating the coefficients of dqi , dpi and dt, one has q˙i p˙i −

∂L ∂t

∂H ∂pi ∂H = − ∂qi ∂H = ∂t =

(35)

The first two equations are the Hamiltonian equations of motion. The Hamiltonian equations are two sets of first order differential equations, rather than the one set of second order differential equations given by the Lagrangian equations of motion. An example is given by motion in one dimension where L =

m 2 x˙ − V (x) 2

(36)

the momentum is given by ∂L ∂ x˙ = m x˙

p =

(37)

Then, the Hamiltonian becomes H

= p x˙ − L m 2 = p x˙ − x˙ + V (x) 2 p2 = + V (x) 2m 20

(38)

which is the same as the energy. For a single particle moving in a central potential, we find that the Hamiltonian in spherical polar coordinates has the form p2ϕ p2θ p2r + + + V (r) 2m 2 m r2 2 m r2 sin2 θ which is the energy of the particle in spherical polar coordinates. H =

(39)

——————————————————————————————————

1.2.2

Exercise 3

Find the Hamiltonian and the Hamiltonian equations of motion for a particle in spherical polar coordinates. ——————————————————————————————————

1.2.3

Solution 3

Using the expression for the Lagrangian   2  2  2  m dr dθ dϕ − V (r) L = + r2 + r2 sin2 θ 2 dt dt dt

(40)

one finds the generalized momenta pr

=



=



=

∂L = m r˙ ∂ r˙ ∂L = m r2 θ˙ ∂ θ˙ ∂L = m r2 sin2 θ ϕ˙ ∂ ϕ˙

(41)

The Hamiltonian is given by H = pr r˙ + pθ θ˙ + pϕ ϕ˙ − L

(42)

which on eliminating r, ˙ θ˙ and ϕ˙ in terms of the generalized momenta, leads to p2ϕ p2r p2θ + V (r) + + 2m 2 m r2 2 m r2 sin2 θ The equations of motion become pr r˙ = m pθ ˙θ = m r2 pϕ ϕ˙ = 2 m r sin2 θ H =

21

(43)

(44)

and  p2ϕ ∂V p2θ = − + − 2 3 3 mr ∂r m sin θ r p2ϕ ∂V = − cos θ − ∂θ m sin3 θ r2 ∂V = − ∂ϕ 

p˙r p˙θ p˙ϕ

(45)

——————————————————————————————————

1.2.4

Time Evolution of a Physical Quantity

Given any physical quantity A then it can be represented by a function of the all the coordinates and momenta, and perhaps explicitly on time t, but not on derivatives with respect to time. This quantity A is denoted by A(qi , pi , t). The rate of change of A with respect to time is given by the total derivative,  X  ∂A ∂A ∂A dA = q˙i + p˙i + (46) dt ∂q ∂p ∂t i i i where the first two terms originate from the dynamics of the particle’s trajectory, the last term originates from the explicit time dependence of the quantity A. On substituting the Hamiltonian equations of motion into the total derivative, and eliminating the rate of change of the coordinates and momenta, one finds  X  ∂A ∂H dA ∂A ∂H ∂A = − + (47) dt ∂q ∂p ∂p ∂q ∂t i i i i i

1.2.5

Poisson Brackets

The Poisson Brackets of two quantities, A and B, is given by the expression  X  ∂A ∂B ∂A ∂B [ A , B ]P B = − (48) ∂qi ∂pi ∂pi ∂qi i The equation of motion for A can be written in terms of the Poisson Bracket, dA ∂A = [ A , H ]P B + dt ∂t

22

(49)

From the definition, it can be seen that the Poisson Bracket is anti-symmetric [ A , B ]P B = − [ B , A ]P B

(50)

The Poisson Bracket of a quantity with itself is identically zero [ A , A ]P B = 0

(51)

If we apply this to the Hamiltonian we find dH dt dH dt

=

[ H , H ]P B +

=

∂H ∂t

∂H ∂t (52)

Thus, if the Hamiltonian doesn’t explicitly depend on time, the Hamiltonian is constant. That is, the energy is conserved. Likewise, if A doesn’t explicitly depend on time and if the Poisson Bracket between H and A is zero, [ A , H ]P B = 0 (53) one finds that A is also a constant of motion dA dt

=

[ A , H ]P B

=

0

(54)

Another important Poisson Bracket relation is the Poisson Bracket of the canonically conjugate coordinates and momenta, which is given by  X  ∂pj ∂qj 0 ∂pj ∂qj 0 [ pj , qj 0 ]P B = − ∂qi ∂pi ∂pi ∂qi i X = 0 − δi,j δi,j 0 i

= − δj,j 0

(55)

The first term is zero as q and p are independent. The last term involves the Kronecker delta function. The Kronecker delta function is given by δi,j = 1 if i = j δi,j = 0 if i 6= j

(56)

and is zero unless i = j, where it is unity. Thus, the Poisson Bracket between a generalized coordinate and its conjugate generalized momentum is − 1, [ pj , qj 0 ]P B = − δj,j 0 23

(57)

while the Poisson Bracket between a coordinate and the momentum conjugate to a different coordinate is zero. We can also show that [ pj , pj 0 ]P B = [ qj , qj 0 ]P B = 0

(58)

These Poisson Brackets shall play an important role in quantum mechanics, and are related to the commutation relations of canonically conjugate coordinate and momentum operators.

24

1.3

A Charged Particle in an Electromagnetic Field

In the classical approximation, a particle of charge q in an electromagnetic field represented by E(r, t) and B(r, t) is subjected to a Lorentz force   1 F = q E(r, t) + r˙ ∧ B(r, t) (59) c The Lorentz force acts as a definition of the electric and magnetic fields, E(r, t) and B(r, t) respectively. In classical mechanics, the fields are observable through the forces they exert on a charged particle. In general, quantum mechanics is couched in the language of potentials instead of forces, therefore, we shall be replacing the electromagnetic fields by the scalar and vector potentials. The electromagnetic fields are solutions of Maxwell’s equations which not only express the fields in terms of the sources and but also form consistency conditions. The scalar and vector potentials simplify Maxwell’s equations since they automatically satisfy the consistency conditions. However, as their definitions specify that they are solutions of firstorder partial differential equations, they are not unique. Despite the ambiguity in the potentials, the physical results that can be derived from them are unique.

1.3.1

The Electromagnetic Field

The electromagnetic field satisfies Maxwell’s eqns., ∇ . B(r, t) = 0 1 ∂ ∇ ∧ E(r, t) + B(r, t) = 0 c ∂t ∇ . E(r, t) = ρ(r, t) 1 ∂ 1 ∇ ∧ B(r, t) − E(r, t) = j(r, t) c ∂t c

(60)

where ρ(r, t) and j(r, t) are the charge and current densities. The last two equations describe the relation between the fields and the sources. The first two equations are the source free equations and are automatically satisfied if one introduced a scalar φ(r, t) and a vector potential A(r, t) such that B(r, t)

= ∇ ∧ A(r, t)

E(r, t)

= − ∇ φ(r, t) −

1 ∂ A(r, t) c ∂t

(61)

In classical mechanics, the electric and magnetic induction fields, E and B, are regarded as the physically measurable fields, and the scalar φ(r, t) and a vector potential A(r, t) are not physically measurable. There is an arbitrariness in the values of the potentials φ(r, t) and A(r, t) as they are defined to be the solutions of differential equations which relate them to the physically measurable 25

E(r, t) and B(r, t) fields. This arbitrariness is formalized in the concept of a gauge transformation, which means that the potentials are not unique and if one replaces the potentials by new values which involve derivatives of any arbitrary scalar function Λ(r, t) 1 ∂ Λ(r, t) c ∂t A(r, t) → A(r, t) + ∇ Λ(r, t) φ(r, t) → φ(r, t) −

(62)

the physical fields, E(r, t) and B(r, t), remain the same. This transformation is called a gauge transformation. Although the laws of physics are formulated in terms of the gauge fields φ(r, t) and A(r, t), the physical results are gaugeinvariant.

1.3.2

The Lagrangian for a Classical Charged Particle

The Lagrangian for a classical particle in an electromagnetic field is expressed as s  r˙ 2 q − q φ(r, t) + A(r, t) . r˙ (63) L = − m c2 1 − 2 c c The canonical momentum now involves a component originating from the field as well as the mechanical momentum p = s

m r˙ 1 −

+ r˙ 2 c2



q A(r, t) c

(64)

The Lagrangian equations of motion for the classical particle are     q q 1 d + A(r, t) = − q ∇ φ(r, t) + ∇ r˙ . A(r, t) (65) m r˙ q c c dt r˙ 2 1 − c2 where the time derivative is a total derivative. The total derivative of the vector potential term is written as ∂ d A(r, t) + r˙ . ∇ A(r, t) A(r, t) = ∂t dt

(66)

as it relates the change of vector potential experienced by a moving particle. The change of vector potential may occur due to an explicit time dependence of A(r, t) at a fixed position, or may occur due to the particle moving to a new position in a non-uniform field A(r, t). ——————————————————————————————————

26

1.3.3

Exercise 4

Show that these equations reduce to the relativistic version of the equations of motion with the Lorentz force law     1 1 d m r˙ q = q E(r, t) + r ˙ ∧ B(r, t) (67) dt c r˙ 2 1 − c2

——————————————————————————————————

1.3.4

Solution 4

The Euler-Lagrange equation is of the form     q q d 1 + A(r, t) = − q ∇ φ(r, t) + ∇ r ˙ . A(r, t) (68) m r˙ q c c dt r˙ 2 1 − c2 On substituting the expression for the total derivative of the vector potential ∂ d A(r, t) = A(r, t) + r˙ . ∇ A(r, t) dt ∂t one obtains the equation  1 d m r˙ q dt 1 −

(69)



q ∂ A(r, t) c ∂t     q q + ∇ r˙ . A(r, t) − r˙ . ∇ A(r, t) c c r˙ 2 c2

= − q ∇ φ(r, t) −

The last two terms can be combined to yield   d q ∂ 1 A(r, t) m r˙ q = − q ∇ φ(r, t) − 2 c ∂t dt r˙ 1 − c2   q ∇ ∧ A(r, t) + r˙ ∧ c

(70)

(71)

The first two terms on the right hand side are identifiable as the expression for the electric field E(r, t) and the last term involving the curl A(r, t) is recognized as involving the magnetic induction field B(r, t) and, therefore, comprises the magnetic component of the Lorentz Force Law. ——————————————————————————————————

27

Despite the fact that the Lagrangian of a charged particle depends on the gauge fields and, therefore, changes form under a gauge transformation, the classical equations of motion are gauge invariant. The equations of motion are gauge invariant since they only dependent on the electromagnetic fields E(r, t) and B(r, t). The gauge invariance of the equations of motion can be seen in a different way, that is by directly applying a gauge transformation to the action S. Under a gauge transformation 1 ∂ Λ(r, t) c ∂t A(r, t) → A(r, t) + ∇ Λ(r, t) φ(r, t) → φ(r, t) −

(72)

the Lagrangian L of a particle with charge q L =

m 2 q r˙ − q φ(r, t) + A(r, t) . r˙ 2 c

(73)

transforms to L → L0 = L +

q c



∂ Λ(r, t) + ∇ Λ(r, t) . r˙ ∂t

 (74)

However, the term proportional to q is recognized as being a total derivative with respect to time. That is, if one considers r to have the time-dependence r(t) of any trajectory connecting the initial and final position of the particle, then q d L → L0 = L + Λ(r(t), t) (75) c dt where the derivative is evaluated over a trajectory of the particle. Hence, under a gauge transformation, the action changes from S to S 0 , where Z tf Z tf q d L(r, r) ˙ + dt Λ(r(t), t) S0 = c ti dt ti   q = S + Λ(rf , tf ) − Λ(ri , ti ) (76) c which only depends on the end-points and not on the path that was chosen. Thus, the gauge transformation only adds a constant term to the action and the trajectory identified by extremal principle is unaffected by the gauge transformation.

1.3.5

The Hamiltonian of a Classical Charged Particle

The Hamiltonian for a charged particle is found from H = p . r˙ − L

28

(77)

which, with the relation between the momentum and the mechanical momentum, q 1 A(r, t) = m r˙ q (78) p − c r˙ 2 1 − c2 leads to s H =

 c2

p −

q A(r, t) c

2

 + m2 c4

+ q φ(r, t)

(79)

The presence of the electromagnetic field results in the two replacements q p → p − A(r, t) c H → H − q φ(r, t) (80) An electromagnetic field is often incorporated in a Hamiltonian describing free charged particles through these two replacements. These replacements retain relativistic invariance as both the pairs E and p and φ(r, t) and A(r, t) form four vectors. This procedure of including an electromagnetic field is based on what is called the minimal coupling assumption. The non-relativistic limit of 2 the Hamiltonian is found by expanding the square root in powers of mp2 c2 and neglecting the rest mass energy m c2 results in the expression  2 q 1 p − A(r, t) H = + q φ(r, t) (81) 2m c which forms the basis of the Hamiltonian used in the Schr¨odinger equation. ——————————————————————————————————

1.3.6

Exercise 5

Derive the Hamiltonian for a charged particle in an electromagnetic field. ——————————————————————————————————

1.3.7

Solution 5

The Lagrangian is given by s L = − m c2

1 −

r˙ 2 q r˙ . A(r, t) − q φ(r, t) + c c2

(82)

so the generalized momentum p is given by p −

m r˙ q A(r, t) = q c 1 − 29

r˙ 2 c2

(83)

and so on inverting this one finds  p −

r˙ =s c

1 −

 A(r, t)

 m2 c2 +

and

q c

r˙ 2 = c

p −

q c

2 A(r, t)

m2 c2  m2

c2

+

p −

(84)

q c

2

(85)

A(r, t)

The Hamiltonian is then found from the Legendre transformation H

= p . r˙ − L s   q r˙ 2 2 = p − A(r, t) . r˙ + m c 1 − 2 + q φ(r, t) c c s  2 q = c m2 c2 + p − A(r, t) + q φ(r, t) c

(86)

—————————————————————————————————— On expanding the quadratic kinetic energy term, one finds the Hamiltonian has the form of a sum of the unperturbed Hamiltonian and an interaction Hamiltonian Hint ,   2 p H = + q φ(r, t) + Hint (87) 2m The interaction Hint couples the particle to the vector potential   q q2 Hint = − p . A(r, t) + A(r, t) . p + A2 (r, t) 2mc 2 m c2

(88)

where the first term linear in A is the paramagnetic coupling and the last term quadratic in A2 is known as the diamagnetic interaction. For a uniform static magnetic field B(r, t) = B, one possible form of the vector potential is A(r) = −

1 r ∧ B 2

The interaction Hamiltonian has the form     2 q q2 Hint = + r ∧ B .p + r ∧ B 2mc 8 m c2 30

(89)

(90)

The first term can be written as the ordinary Zeeman interaction between the orbital magnetic moment and the magnetic field    2 q q2 Hint = − B.L + r ∧ B (91) 2mc 8 m c2 where the orbital magnetic moment M is related to the orbital angular momentum via q M = + L (92) 2mc Hence, the ordinary Zeeman interaction has the form of a dipole interaction

L

M = q / (2mc) L

q

Figure 6: A particle with charge q and angular momentum L, posseses a magq netic dipole moment M given by M = + 2 m c L. with the magnetic field HZeeman = − M . B

(93)

which has the tendency of aligning the magnetic moment parallel to the field. Elementary particles such as electrons have another form of magnetic moment and angular momentum which is intrinsic to the particle, and is not connected to any physical motion of the particle. The intrinsic angular momentum of elementary particles is known as spin. There are two general approaches that can be taken to Quantum Mechanics. One is the path integral approach which was first developed by Dirac and then 31

popularized by Feynmann. This approach is based on the use of the Lagrangian formulation of classical mechanics. The other approach which is more common, and is the one that we shall follow exclusively, is based on the Hamiltonian formulation of classical mechanics.

32

2

Failure of Classical Mechanics

Classical Mechanics fails to correctly describe some physical phenomena. This first became apparent at the atomic level. Historically, the failure of classical mechanics was first manifested after Rutherford’s discovery of the structure of the atom. Classically, an electron orbiting around a charged nucleus should continuously radiate energy according to Maxwell’s Electromagnetic Theory. The radiation leads to the electron experiencing a loss of energy, and thereby continuously reducing the radius of the electrons orbit. Thus, the atom becomes unstable as the electron spirals into the nucleus. However, it is a well established experimental fact that atoms are stable and that the atomic energy levels have discrete values for the energy. The quantization of the energy levels is seen through the Franck-Hertz experiment, which involves inelastic collisions between atoms. Other experimental evidence for the quantization of atomic energy levels is given by the emission and absorption of electromagnetic radiation. For example, the series of dark lines seen in the transmitted spectrum when light, with a continuous spectrum of wavelengths, falls incident on hydrogen gas is evidence that the excitation spectrum consists of discrete energies. Niels Bohr discovered that the excitation energy ∆E and the angular frequency of the light ω are related via ∆E = h ¯ ω (94) The quantity ¯h is known as Planck’s constant and has the value of h ¯

= =

1.0545 × 10−34 J s 0.65829 × 10−15 eV s

(95)

The Balmer, Lyman and Paschen series of electromagnetic absorption by hydrogen atoms establishes that ∆E takes on discrete values. Another step in the development of quantum mechanics occurred when Louis de Broglie postulated wave particle duality, namely that entities which have the attributes of particles also posses attributes of waves. This is formalized by the relationships E p

= h ¯ ω = h ¯ k

(96)

The de Broglie relations were verified by Davisson and Germer in their experiments in which a beam of electrons were placed incident on the surface of a crystalline solid, and the reflected beam showed a diffraction pattern indicative of the fact that the electrons have a wave length λ. The diffraction condition relates the angle of the diffracted beam to the ratio of the separation between the planes of atoms and the wavelength λ. Furthermore, the diffraction condition showed a variation with the particle’s energies which is consistent with the wavelength momentum relation.

33

2.1

Semi-Classical Quantization

The first insight into quantum phenomena came from Niels Bohr, who imposed an additional condition on Classical Mechanics1 . This condition, when imposed on systems where particles undergo periodic orbits, reduces the continuous values of allowed energies to a set of discrete energies. This semi-classical quantization condition is written as I pi dqi = ni h (97) where pi and qi are the canonically conjugate momentum and coordinates of Lagrangian or Hamiltonian Mechanics, and ni is an integer from the set (0,1,2,3,4,. . . ,∞) and h is a universal constant. The integration is over one period of the particle’s orbit in phase space2 . The discrete values of the excitation energies found for the hydrogen atom produces reasonably good agreement with the excitation energies found for the absorption or emission of light from hydrogen gas. ——————————————————————————————————

2.1.1

Exercise 6

Assuming that electrons move in circular orbits in the Coulomb potential due to a positively charged nucleus, (of charge Z e ), find the allowed values of the energy when the semi-classical quantization condition is imposed. ——————————————————————————————————

2.1.2

Solution 6

The Lagrangian L is given in terms of the kinetic energy T and the scalar electrostatic potential V by L = T − V . Then in spherical polar coordinates (r, θ, ϕ), one finds the Lagrangian L =

m r˙ 2 m r2 θ˙2 m r2 sin2 θ ϕ˙ 2 Z e2 + + + 2 2 2 r

(98)

The Euler-Lagrange equation for the radius r is given by ∂L ∂r 1 N.

=

  d ∂L dt ∂ r˙

Bohr, Phil. Mag. 26, 1 (1913). showed that this quantization rule can be applied to certain non-separable systems [A. Einstein, Deutsche Physikalische Gesellschaft Verhandlungen 19, 82 (1917)]. In a non-ergodic system, the accessible phase space defines a d-dimensional torus. The quantization condition can be applied to any of the d independent closed loops on the torus. These loops do not have to coincide with the system’s trajectory. 2 Einstein

34

Z e2 m r θ˙2 + m r sin2 θ ϕ˙ 2 − = m r¨ r2 while the equation of motion for the polar angle θ is given by   ∂L d ∂L = ∂θ dt ∂ θ˙   2 d mr sin 2θ ϕ˙ 2 = m r2 θ˙ 2 dt

(99)

(100)

and finally we find   d ∂L = dt ∂ ϕ˙   d 2 2 0 = m r sin θ ϕ˙ dt

∂L ∂ϕ

(101)

According to the assumption, the motion is circular, which we choose to be in the equatorial plane (r = a, θ = π2 ). Thus, one has Z e2 = m a ϕ˙ 2 a2 0 = m a2 ϕ¨

(102)

Hence, we find that the angular velocity is a constant, ϕ˙ = ω. To impose the semi-classical quantization condition we need the canonical momentum. The canonical momenta are given pr

=



=



=

∂L = 0 ∂ r˙ ∂L = 0 ∂ θ˙ ∂L = m r2 sin2 θ ϕ˙ = m a2 ω ∂ ϕ˙

The semi-classical quantization condition becomes I pϕ dϕ = m a2 ω 2π = nϕ h

(103)

(104)

From the above one has the two equations m a2 ω

= nϕ ¯h

2

Z e = m a ω2 (105) a2 On solving these for the Bohr radius a and angular frequency ω, one finds ω = a =

Z 2 e4 n3ϕ ¯h3 n2ϕ ¯h2 Z e2 m 35

(106)

Substituting these equations in the expression for the energy E, one finds the expression first found by Bohr E E E

Z e2 m a2 ω 2 − 2 a Z e2 = − 2a m Z 2 e4 = − 2 n2ϕ ¯h2

=

(107)

This agrees with the exact (non-relativistic) quantum mechanical expression for the energy levels of electrons bound to a H ion. ——————————————————————————————————

2.1.3

Exercise 7

Find the energy of a one-dimensional simple Harmonic oscillator, of mass m and frequency ω, when the semi-classical quantization condition is imposed. ——————————————————————————————————

2.1.4

Solution 7

The Hamiltonian is given by H(p, q) =

p2 m ω2 q2 + 2m 2

(108)

Hamilton’s equations of motions are ∂H = − m ω2 q ∂q ∂H p + = ∂p m



= −



=

(109)

Thus, we have the equation of motion q¨ + ω 2 q = 0

(110)

which has the solution in the form q(t) p(t)

= A sin(ωt + φ) = m ω A cos(ωt + φ)

36

(111)

The semi-classical quantization condition becomes I Z 2π p(t) dq(t) = m ω A2 dϕ cos2 ϕ 0

= m ω A2 2 π = n h Thus, we find A2 =

2 n¯ h m ω

where h ¯ =

E = H(p, q) =

h 2π

(112)

and the energy becomes

m ω 2 A2 = n ¯h ω 2

(113)

This should be compared with the exact quantum mechanical result E = ¯hω ( n + 12 ). The difference between these results become negligible for large n. That is for a fixed Energy E, the results approach each other in the limit of large n or equivalently for small h ¯ . This is example is illustrative of the correspondence principle, which states that Quantum Mechanics, should closely approximate Classical Mechanics where Classical Mechanics is known to provide an accurate description of nature, and that this in this limit ¯h can be considered to be small. ——————————————————————————————————

37

3

Principles of Quantum Mechanics

The Schr¨ odinger approach to Quantum Mechanics is known as wave mechanics. The Schr¨ odinger formulation is in terms of states of a system which are represented by complex functions defined in Euclidean space, Ψ(r) known as wave functions and measurements are represented by linear differential operators. The Schr¨ odinger approach, though most common is not unique. An alternate approach was pursued by Heisenberg, in which the state of the system are represented by column matrices and measurements are represented by square matrices. This second approach is known as Heisenberg’s matrix mechanics. These two approaches were shown to be equivalent by Dirac, who developed an abstract formulation of Quantum Mechanics using an abstract representation of states and operators. We shall first consider the system at a fixed time t, say t = 0. The wave function, Ψ(r), represents a state of a single particle at that instant of time, and has a probabilistic interpretation. Consider an ensemble of N identical and non-interacting systems, each of which contains a measurement apparatus and a single particle. Each measurement apparatus has its own internal reference frame with its own origin. A measurement of the position r (referenced to the coordinate system attached to the measurement apparatus) is to be made on each particle in the ensemble. Just before the measurements are made, each particle is in a state represented by Ψ(r). Measurements of the positions of each particle in the ensemble will result in a set of values of r that represents points in space, referenced w.r.t. the internal coordinate system3 . The probability that a measurement of the position r of a particle will give a value in the infinitesimal volume d3 r containing the point r, is given by P (r) d3 r = | Ψ(r) |2 d3 r

(114)

Thus, P (r) d3 r is the probability of finding the particle in the infinitesimal volume d3 r located at r. The probability is directly proportional to the size of the volume d3 r, and P (r) = | Ψ(r) |2 is the probability density. Since the particle is somewhere in three-dimensional space, the probability is normalized such that Z (115) | Ψ(r) |2 d3 r = 1 This normalization condition will have to be enforced on the wave function if it is to represent a single-particle state. The normalization condition must be true for all times, if the particle number is conserved4 . 3 Alternatively, instead of considering an ensemble of identical systems, one could consider performing N successive measurements on a single system. However, before each successive measurement is made, one would have to reset the initial condition. That is, the system should be prepared so that, just before each measurement is made, the particle is in the state described by Ψ(r). 4 For well-behaved functions, the normalization condition implies that | Ψ(r) |2 vanishes as |r| → ∞.

38

Given two states, Ψ(r) and Φ(r), one can define an inner product or overlap matrix element as the complex constant given by Z d3 r Φ∗ (r) Ψ(r) (116) where the integration runs over all volume of three-dimensional space. It should be noted that inner product of two wave functions depends on the order that the wave functions are specified. If the inner product is taken in the opposite order, one finds  Z ∗ Z d3 r Ψ∗ (r) Φ(r) = d3 r Φ∗ (r) Ψ(r) (117) which is the complex conjugate of the original inner product. The normalization of the wave function Ψ(r) just consists of the inner product of the wave function with itself. The interpretation of the squared modulus of the wave function as a probability density requires that the normalization of a state is unity. We should note that if all observable quantities for a state always involve the wave function times its complex conjugate, then the absolute phase of the wave function is not observable. Only phase differences are measurable. Therefore, it is always possible to transform a wave function by changing its phase   i Λ(r) 0 Ψ(r) → Ψ (r) = exp Ψ(r) (118) ¯h where Λ(r) is an arbitrary real function.

39

3.1

The Principle of Linear Superposition

In quantum mechanics, a measurement of a physical quantity of a single particle which is in a unique state will result in a value of the measured quantity that is one of a set of possible results an . Repetition of the measurement on a particle in exactly the same initial state may yield other values of the measured quantity (such as am ). The probability distribution for the various results an is governed by the particular state of the system that the measurement is being performed on. This suggests that a state of a quantum mechanical system, at any instant, can be represented as a superposition of states corresponding to the different possible results of the measurement. Let Φn (r) be a state such that a measurement of A on the state will definitely give the result an . The simplest way of making a superposition of states is by linearly adding multiples of the wave functions Φn (r) corresponding to the possible results. Thus, the principle of linear superposition can be stated as X Ψ(r) = Cn Φn (r)

(119)

n

where the expansion coefficients Cn are complex numbers. The expansion coefficients can be determined from a knowledge of the wave function of the state Ψ(r) and the set of functions, Φn (r), representing the states in which a measurement of A is known to yield the result an . The expansion coefficients Cn are related to the probability that the measurement on the state Ψ(r) results in the value an . The principle of linear superposition of the wave function can lead to interference in the results of measurements. For example, the results of a measurement of the position of the particle r, leads to a probability distribution P (r) according to P (r) d3 r

= | Ψ(r) |2 d3 r X ∗ = Cm Cn Φ∗m (r) Φn (r) d3 r

(120)

n,m

On isolating the terms with n = m, one finds terms in which the phases of the wave functions Φn (r) and the phases of the complex numbers Cn separately cancel. The remaining terms, in which n 6= m, represent the interference terms. Thus X X ∗ P (r) d3 r = | Cn |2 | Φn (r) |2 + Cm Cn Φ∗m (r) Φn (r) d3 r (121) n

n6=m

The coefficient | Cn |2 is the probability that the state represented by Ψ(r) is in the state n.

40

As an example, consider a state which is in a superposition of two states each of which represents a state of definite momentum, p = ¯h k and p = − ¯h k. The forward and backward travelling states are   Φk (r) = Ck exp + i k . r   Φ−k (r) = C−k exp − i k . r (122) These states are not normalizable and, therefore, each state must represent beams of particles with definite momentum in Rwhich the particles are uniformly distributed over all space. Since the integral | Ψ(r) |2 d3 r → ∞, the beam must be considered to contain an infinite number of particles. The probability densities or intensities of the two independent beams are given by Pk (r) P−k (r)

= | Φk (r) |2 = | Ck |2 = | Φ−k (r) |2 = | C−k |2

(123)

We shall assume that these beams have the same intensities, that is | C−k |2 = | C+k |2 . Then, in this case the wave function can be expressed in terms of the phases of C± = | C | exp[ i δ± ]. When the beams are superimposed, the state is described by the wave function   (δ+ + δ− ) Ψ(r) = | C | exp i 2    ! (δ+ − δ− ) (δ+ − δ− ) × exp + i (k . r + ) + exp − i (k . r + ) 2 2 (124) This state is a linear superposition and has a probability density for finding the particle at r given by P (r) where   (δ+ − δ− ) k.r + (125) P (r) = 4 | C |2 cos2 2 Thus, the backward and forward travelling beam interfere, the superposition gives rise to consecutive planes of maxima and minima. The maxima are located at (δ+ − δ− ) = πn 2

(126)

(δ+ − δ− ) π = (2n + 1) 2 2

(127)

k.r + and the minima are located at k.r +

41

There is a strong analogy between Quantum Mechanics and Optics. The wave function Ψ(r, t) plays the role corresponding to the electric field E(r, t). Both the wave function and the electric field obey the principle of linear superposition. The probability density of finding the particle at point r, | Ψ(r, t) |2 plays the role of the intensity of light, I, which is proportional to | E(r, t) |2 . In fact, the intensity of light is just proportional to the probability density of finding a photon at the point r. The phenomenon of interference occurs in Quantum Mechanics and also in Optics. The analogy between Quantum Mechanics and Optics is not accidental, as Maxwell’s equations are intimately related to the ”Schr¨ odinger equation” for a massless particle with intrinsic spin S = 1.

42

3.2

Wave Packets

By combining momentum states with many different values of the momentum, one can obtain states in which the particle is essentially localized in a finite volume. These localized states are defined to be wave packets. Wave packets are the closest one can get to a classical state of a free particle, which has a well defined position and momentum. For a quantum mechanical wave packet, the distribution of results for measurements of the position and momentum are sharply peaked around the classical values. The wave function can be expressed

1

Re Im

Ψ(x) Ψ( )

0.5

0 -3

-2

-1

0

1

2

3

-0.5

-1

x

Figure 7: The real and imaginary parts of a wavefunction Ψ(x) representing a Gaussian wave packet in one dimension. as a Fourier transform  Ψ(r) =

1 2π

 32 Z

3

d k Φ(k) exp



 + ik.r

(128)

where Φ(k) is related to the momentum probability distribution function5 . Since the exponential factor represents states of different momenta, this relation is an example of how a wave function can be expanded in terms of states corresponding to the various possible results of a physical measurement. The momentum 5 This is an example of the principle of linear superposition, in which the discrete variable n has been replaced by a Riemann sum over the continuous variable k and the expansion coefficients Cn are proportional to Φ(k).

43

probability distribution function is related to the wave vector probability distribution function Pk (k), defined by Pk (k) d3 k = | Φ(k) |2 d3 k

(129)

The function Φ(k) can be determined from knowledge of Ψ(r) from the inverse relation  3 Z   1 2 3 Φ(k) = d r Ψ(r) exp − i k . r (130) 2π These results are from the mathematical theory of Fourier Transformations. The consistency of the Fourier Transform with the Inverse Fourier Transform can be seen by combining them via  Ψ(r)

=  =

1 2π

 23 Z

1 2π

3 Z

d3 k Φ(k) exp 3

Z

d k

3 0



 + ik.r 

0

d r Ψ(r ) exp

0

ik.(r − r )

 (131)

together with the representation of the three-dimensional Dirac delta function 3

0

δ (r − r ) =



1 2π

3 Z

3

d k exp



0

ik.(r − r )

 (132)

On inserting the integral representation of the Dirac delta function eqn(132) into eqn(131), one finds an equation which is the formal definition of the Dirac delta function Z d3 r0 Ψ(r0 ) δ 3 ( r − r0 ) (133) Ψ(r) =

That equation (132) provides a representation of the three-dimensional Dirac delta function can by seen directly by factorizing it into the product of three independent one-dimensional delta functions as δ 3 ( r − r0 ) = δ( x − x0 ) δ( y − y 0 ) δ( z − z 0 )

(134)

and then comparing with the right hand side which can also be factorized into three independent one-dimensional integrals. Each factor in the threedimensional delta function of eqn(132) can be replaced by the representation of the one-dimensional delta function as a limit of a sequence of Lorentzian functions of width , δ( x − x0 ) =

lim

 → 0

1  0 π ( x − x )2 + 2

(135)

The sequence of function is shown in fig(8). Then on evaluating the integrations over each of the one-dimensional variables in the right hand side of eqn(132) 44

Dirac delta dunction δε(x)

2

δε(x) = 1/π ε/(ε 2+x2)

1.5

lim ε → 0

1

0.5

0 -2

-1.5

-1

-0.5

0

0.5

1

1.5

2

x

Figure 8: The Dirac delta function δ(x). The Dirac delta function is defined as the limit,  → 0 of a sequence of functions δ (x). and using the replacement ( x − x0 ) → ( x − x0 ) ± i  needed to keep the integral convergent, one finds that each of the three factors have the same form   Z +L 1 = lim dk exp i k ( x − x0 ) 2 π L→∞ −L  Z +L   1 0 = lim dk exp i k ( x − x ) −  k 2 π L→∞ 0   Z 0 + dk exp i k ( x − x0 ) +  k −L   1 1 1 = − + 2π i ( x − x0 ) −  i ( x − x0 ) +  1  = lim (136)  → 0 π ( x − x0 )2 + 2 which equals the corresponding representation of the delta function on the left hand side. This completes the identification, and proves that the inverse Fourier transform of the Fourier transform is the original function. It has also proved that any square integrable function, i.e. normalizable function, can be expanded as a sum of momentum eigenfunctions. The momentum eigenstates interfere destructively almost everywhere, except at the position where the wave packet is

45

peaked. In general, the d-dimensional momentum distribution is related to the distribution of k vectors, | Φ(k) |2 by the relation 2 Z Pp (p) = dd k δ d ( p − ¯h k ) Φ(k)   p 1 Φ d ¯ h ¯h

=

2

(137) d

In future, we shall find it convenient to include factors of h ¯ − 2 into the definition of the d-dimensional Fourier transform so that, on squaring the modulus, they give the properly normalized momentum distribution function. ——————————————————————————————————

3.2.1

Exercise 8

Given a wave function Ψ(x) where     ( x − x0 )2 exp + i k0 x Ψ(x) = C exp − 4 δx2

(138)

determine C, up to an arbitrary phase factor, and determine Φ(kx ). Later, we shall see that | Φ(kx ) |2 is proportional to the probability distribution of the x component of momentum of the system. In three dimensions | Φ(k) |2 d3 k is the probability of finding the system with a wave vector k in an infinitesimal volume d3 k around the point k. ——————————————————————————————————

3.2.2

Solution 8

The magnitude of the coefficient C is determined from the normalization condition Z +∞ dx | Ψ(x) |2 = 1 (139) − ∞

which is evaluated as | C |2

Z

+ ∞

 dx exp



− ∞

( x − x0 )2 2 δx2

 = 1

(140)

On changing the variable of integration from x to y where y =

x − x0 √ 2 δx 46

(141)

the normalization condition becomes   Z +∞ √ 2 2 1 = |C | 2 δx dy exp − y − ∞ √ 1 = | C |2 2 π δx

(142)

Hence, the magnitude of C is found as |C | =

1 1

(143)

1

(2 π) 4 δx 2

Real-space distribution function 0.5

l ψ(x) l2

0.4

0.3

0.2

x0 0.1

0 -2

-1.5

-1

-0.5

0

0.5

1

1.5

2

x

Figure 9: The real space distribution function | Ψ(x) |2 . The distribution function is centered around x0 . The k-space wave function is found from Z +∞ 1 Φ(kx ) = √ dx Ψ(x) exp[ − i kx x ] 2π −∞   Z +∞ 1 ( x − x0 )2 = dx exp − exp[ + i k0 x ] exp[ − i kx x ] 3 1 4 δx2 (2 π) 4 δx 2 − ∞ (144) Combining the exponential factors and completing the square by changing variable from x to z where z = x − x0 + 2 i ( kx − k0 ) δx2 47

(145)

one finds that Φ(kx )

=

+ ∞

 ( z )2 3 1 4 δx2 (2 π) 4 δx 2 − ∞     × exp − ( k0 − kx )2 δx2 exp − i ( kx − k0 ) x0 Z

1

1

= δx 2



2 π





dz exp

 14

 exp

− ( k0 − kx )2 δx2



 exp

 − i ( kx − k0 ) x0 (146)

Thus, we see that in kx space the modulus squared wave function | Φ(kx ) |2 is centered around k0 and has a k width which is proportional to δx−1 .

Momentum-space distribution function 1

l Φ(p) l2

0.8

0.6

0.4

0.2

p0

0 -2

-1.5

-1

-0.5

0

0.5

1

1.5

2

p

Figure 10: The momentum space distribution function | Φ(p) |2 . The momentum distribution is centered about p0 . ——————————————————————————————————

3.2.3

Exercise 9

Given the wave function  Ψ(x) = C exp

48

 − λ|x|

(147)

where λ is a positive real number, find C and the properly normalized momentum distribution function.

Real-space distribution function 1.2 1

l ψ(x) l

2

0.8 0.6 0.4 0.2 0 -2

-1.5

-1

-0.5

0

0.5

1

1.5

2

x

Figure 11: The real space distribution function | Ψ(x) |2 . ——————————————————————————————————

3.2.4

Solution 9

The normalization condition is given by Z +∞ dx | Ψ(x) |2 = 1

(148)

− ∞

which is evaluated as 2

Z

+ ∞

|C |

 dx exp

 − 2λ|x|

= 1

(149)

− ∞

The integration can be broken up into two parts, one over the range ( − ∞ , 0 ) and the second over the range ( 0 , + ∞ ). On replacing | x | by ± x in the appropriate interval, we have  Z 0     Z +∞ 1 = | C |2 dx exp + 2 λ x + dx exp − 2 λ x − ∞

0

49

1

= | C |2



1 1 + 2λ 2λ

 (150)

Thus, the normalization is given by √

|C | =

λ

(151)

Momentum-space distribution function 1

l Φ(p) l2

0.8

0.6

0.4

0.2

0 -2

-1.5

-1

-0.5

0

0.5

1

1.5

2

p

Figure 12: The momentum space distribution function | Φ(p) |2 . The k-space wave function is given by   Z +∞ 1 dx Ψ(x) exp − i kx x Φ(kx ) = √ 2π −∞ r     Z +∞ λ = dx exp − λ | x | exp − i kx x 2π −∞ (152) This is evaluated, as before, by breaking up the integral into two intervals. The first interval is an integration over the range ( − ∞ , 0 ) and the second interval has x in the range ( 0 , + ∞ ). Thus, we find r   λ 1 1 Φ(kx ) = + 2π λ − i kx λ + i kx 50

r =

λ 2λ 2 2 π λ + kx2

(153)

The width of the kx distribution is proportional to λ. The momentum distribution Pp (px ) is given by   px Pp (px ) dpx = Φ ¯h

2 dpx ¯h

(154)

Thus, we find that the momentum distribution function is given by Pp (px ) =

2 ¯ 3 λ3 h π ( ¯h2 λ2 + p2x )2

(155)

which is properly normalized to unity. ——————————————————————————————————

51

3.3

Probability, Mean and Deviations

In Quantum Mechanics one often deals with systems in a well defined state, however, the result of a particular measurement is indeterminate. That is, usually, the result of the measurement cannot be predicted with absolute certainty. Therefore, we shall review some aspects of probabilities and averages.

Sequential measurements of A 2

N=20 1.5

Ai

∆A4

∆Arms

1

0.5

∆A9

0 0

5

10

15

20

i

Figure 13: The set of results of N sequential measurements of A. The average value < A > and the magnitude of the root mean squared deviation ∆Arms of this set of measurements are indicated in red. Also shown are ∆Ai which represent the deviations of some individual data points from the average value. The average or mean value of a measured quantity A, denoted by A, is expressed as an integral (and/or a sum) over all the possible values of the measurement A, weighted with the probability distribution function P (A) via Z A = dA P (A) A (156) The integration runs over all possible (real) values of A. The higher moments Am have averages, Am , which are given by Z m A = dA P (A) Am (157)

52

Probability of measurements of A 0.2

N →∞

P(A)

0.15

0.1

0.05

0 0.1

0.3

0.5

0.7

0.9

1.1

1.3

1.5

1.7

1.9

A

Figure 14: The probability distribution P (A) is defined as the relative frequency of the occurrence of result A, in the limit N → ∞ where N is the number of measurements. The deviation of the variable A from its mean value is defined by ∆A = A − A

(158)

The average value of ∆A is zero, as can be seen from ∆A = A − A = 0

(159)

which just reflects the fact that the variable A fluctuates equally on both sides of the average value A. A measure of the size of the fluctuations can be found from the mean squared deviation ∆A2 via Z ∆A2 = dA P (A) ( A − A )2 Z 2 = dA P (A) ( A2 − 2 A A + A ) 2

= A2 − A

(160)

The mean squared deviation is non-zero as ∆A2 is always positive definite. The variance or root mean squared (r.m.s.) deviation ∆Arms provides a measure of the fluctuations in A. p ∆Arms = ∆A2 53

q ( A − A )2 q 2 = ( A2 − A ) =

(161)

The r.m.s. deviation is a measure of the uncertainty in the value of A. From the examples of wave packets given in exercise 4, it can be found that the uncertainty in the i-th component of the momentum and the i-th component of the particle’s position satisfies the inequality ∆ri

rms

∆pi

rms



¯ h 2

(162)

This is an example of the Heisenberg uncertainty relation, applied to the variables r and p. ——————————————————————————————————

3.3.1

Exercise 10

Given P (x) where  P (x) =

1 2π

 12

δx−1 exp

 −

( x − x0 )2 2 δx2

 (163)

find x, x2 , ∆x2 and ∆x4 . ——————————————————————————————————

3.3.2

Solution 10

The average value of x, denoted as x is given by Z +∞ x = dx P (x) x − ∞ + ∞

Z =

dx | Ψ(x) |2 x

− ∞

=

1 √ 2π

Z

+ ∞ −1

dx δx − ∞

 exp

( x − x0 )2 − 2 δx2

 x (164)

Changing variable to y = x − x0 , one has   Z +∞ 1 y2 x = √ dy δx−1 exp − ( y + x0 ) 2 δx2 2π −∞ (165) 54

The term linear in y is odd and, therefore, vanishes. The term proportional to x0 factors out of the integral to yield   Z +∞ 1 y2 x = x0 √ dy δx−1 exp − 2 δx2 2π −∞ Z +∞ = x0 dx P (x) − ∞

= x0

(166)

as P (x) is normalized to unity. Thus, the average value of x is x0 . The average value of x2 is given by Z +∞ x2 = dx P (x) x2 − ∞ + ∞

Z

dx | Ψ(x) |2 x2

= − ∞

=



1 2π

Z

+ ∞

dx δx−1 exp

− ∞

 −

( x − x0 )2 2 δx2



x2 (167)

Changing variable to y = x − x0 one has   Z +∞ 1 y2 x2 = √ dy δx−1 exp − ( y 2 + 2 y x0 + x20 ) 2 δx2 2π −∞ (168) The term linear in y is odd and, therefore, vanishes. The term proportional to x20 factors out of the integral to yield   Z +∞ 1 y2 x2 = √ dy δx−1 y 2 exp − + 2 δx2 2π −∞   Z +∞ 1 y2 + x20 √ dy δx−1 exp − 2 δx2 2π −∞   Z +∞ 1 y2 −1 2 = √ dy δx y exp − + 2 δx2 2π −∞ Z +∞ + x20 dx P (x) − ∞

= x20 + √

1 2π

Z

+ ∞

dy δx−1 y 2 exp

 −

− ∞

y2 2 δx2

 (169)

as P (x) is normalized to unity. Since we have   Z +∞ 1 1 √ dy exp − α y 2 = √ π −∞ α 55

(170)

then by differentiating with respect to α we find     Z +∞ Z +∞ 1 ∂ 1 √ √ dy exp − α y 2 dy y 2 exp − α y 2 = − ∂α π − ∞ π −∞ ∂ 1 √ = − ∂α α 1 1 √ (171) = + 2 α3 Hence x2 = x20 + δx2

(172)

Thus, the average value of x2 is x20 + δx2 . The mean squared deviation is then found as Z +∞ 2 2 ∆xrms = ∆x = dx P (x) ( x − x0 )2 =

1 √ 2π

Z

− ∞ + ∞

dy δx−1 y 2 exp



− ∞



y2 2 δx2

 (173)

The integration was previously evaluated, and yields ∆x2rms = δx2

(174)

Finally, we can show that Z +∞ 4 ∆x = dx P (x) ( x − x0 )4 − ∞

=

1 √ 2π

Z

+ ∞ −1

dy δx

4

y exp

− ∞



y2 − 2 δx2

 (175)

The integration can be performed by taking the second derivative with respect to α of   Z +∞ 1 1 2 √ dy exp − α y = √ (176) π −∞ α Then, by differentiating with respect to α, we find     Z +∞ Z +∞ 1 ∂2 1 4 2 2 √ √ dy y exp − α y = + dy exp − α y ∂α2 π − ∞ π −∞ ∂2 1 √ ∂α2 α 3 1 √ = + 4 α5

= +

56

(177)

Thus, we have ∆x4

Z

+ ∞

=

dx P (x) ( x − x0 )4

− ∞

= =

1 √ 2π 3 δx4

Z

+ ∞ −1

dy δx



4

y exp

− ∞

y2 − 2 δx2

 (178)

——————————————————————————————————

3.3.3

Exercise 11

Find the r.m.s. deviation in the particle’s position ∆xrms and momentum ∆px rms for the wave packet    1   ( x − x0 )2 1 4 − 12 exp + i k0 x (179) δx exp − Ψ(x) = 4 δx2 2π

——————————————————————————————————

3.3.4

Solution 11

From the results of Exercise 10 we immediately find that ∆xrms = δx. The kx space wave function is given by  1     1 2 4 2 2 2 Φ(kx ) = δx exp − ( k0 − kx ) δx exp − i ( kx − k0 ) x0 π (180) Thus, with px = ¯h kx one has the momentum probability distribution P (px ) given by 2 px  dpx P (px ) dpx = Φ ¯h ¯h  1   2 2 δx 2 2 = exp − 2 ( k0 − kx ) δx dpx (181) π ¯h Hence, this is a Gaussian distribution centered at px = h ¯ k0 , so the average value of the momentum is h ¯ k0 . The width of the Gaussian is = 2 h¯δx , so ∆px

rms

∆xrms = 57

¯ h 2

(182)

which is an example of the Heisenberg uncertainty principle in which the equality holds. ——————————————————————————————————

3.3.5

Exercise 12

Given that a particle moving in one dimension is in a state given by 

1 2 L

 12 if | x | < L

Ψ(x) = 0

otherwise

(183)

find the probability that it is found in a momentum eigenstate, and verify that

0.8

Ψ(x)

0.6

0.4

0.2

0 -2

-1

0

1

2

x/L

Figure 15: The wavefunction Ψ(x) given in Exercise 12. the probability distribution is properly normalized. ——————————————————————————————————

58

3.3.6

Solution 12

The real space wave function is expressed in terms of the momentum space wave function via   Z dp px √ Ψ(x) = exp + i Φ(p) (184) ¯h 2 π ¯h and the inverse relation is Z

dx √ exp 2 π ¯h

Φ(p) =



px − i ¯h

 Ψ(x)

(185)

The properly normalized momentum space wave function Φ(p) is evaluated as + L

  dx 1 px √ exp − i ¯h 2 π ¯h 2 L − L   Z +L 1 px √ dx exp − i = ¯h 4 π L ¯h − L   r sin ph¯L L = p L π ¯h h ¯ Z

Φ(p)



=

(186)

The normalized momentum probability density P (p) is given by the modulus squared of the momentum space wave function 2 P (p) = Φ(p)   p L 2 sin h ¯ L (187) =  2 π ¯h p L h ¯

The distribution P (p) is properly normalized since Z

+ ∞

dx − ∞

sin2 x = π x2

(188)

——————————————————————————————————

59

Momentum Probability Distribution 0.4

P(p) !/L

0.3

0.2

0.1

0 -8

-6

-4

-2

0

2

4

6

8

pL/!

Figure 16: The momentum probability distribution function P (p) for the Ψ(x) given in Exercise 12.

3.4

Operators and Measurements

In Schr¨ odinger’s wave mechanics, a physical or measurable quantity is repreˆ An operator Aˆ is defined by its sented by a (linear) differential operator A. action on the states represented by arbitrary wave functions Ψ(r). The operator Aˆ transforms the state Ψ(r) into another state Φ(r) via Aˆ Ψ(r) = Φ(r)

(189)

A measurement on a quantum mechanical system generally disturbs the system thereby causing a change in the state of the system. As the measurement results in a change in the state of a system, the measurement is represented by an operator. A linear operator Aˆ is defined by its action on a state Ψ(r) that is a linear superposition, X Cn Φn (r) (190) Ψ(r) = n

An operator Aˆ acting on a state Ψ(r) formed as a linear superposition is defined

60

to be a linear operator if, and only if, it produces the result X X Aˆ Cn Φn (r) = Cn Aˆ Φn (r) n

(191)

n

That is, the resulting state must be equivalent to the superposition of the states formed by Aˆ acting on the components Φn (r). The position operator ˆr is given by the vector r. Its operation on the state Ψ(r) is just that of multiplication by r. ˆr Ψ(r) = r Ψ(r)

(192)

The momentum operator p ˆ is given by − i ¯h ∇. The operator ∇ is given in Cartesian coordinates as ∂ ∂ ∂ ∇ = eˆx + eˆy + eˆz (193) ∂x ∂x ∂z where eˆx , eˆy and eˆz are the three orthogonal unit vectors used to define a Cartesian coordinate system. The action of the momentum operator on a state Ψ(r) is given by p ˆ Ψ(r) = − i ¯h ∇Ψ(r) (194)

2

h ¯ The kinetic energy operator is given by Tˆ = − 2m ∇2 . Its action on a state Ψ(r) is given by ¯h2 Tˆ Ψ(r) = − ∇2 Ψ(r) (195) 2m where the Laplacian is given in Cartesian coordinates as

∇2 =

∂2 ∂2 ∂2 + + 2 2 ∂x ∂y ∂z 2

(196)

The operators have no mathematical meaning except when they act upon a wave function.

3.4.1

Operator Equations

Just as in Hamiltonian Mechanics, the operators corresponding to physically measurable quantities can be expressed in terms of the momentum and position operators. This requires compounding operators, via addition and multiplication. The results of compounding operators can lead to operator equations. Operator equations, if valid, must be true if the same expression results when the operators on both side of the equality sign act on the same arbitrary wave function. Usually in writing an operator equation the wave function is suppressed, but it is implicitly assumed to be present.

61

3.4.2

Operator Addition

ˆ each defined by their action on every Given two differential operators Aˆ and B, wave function Ψ(r) by Aˆ Ψ(r) = ΦA (r) ˆ Ψ(r) = ΦB (r) B

(197)

where the final wave functions ΦX (r) are dependent on the particular choice of ˆ ) and the wave function Ψ(r). The sum of two operators is defined as ( Aˆ + B is given by the linear sum ˆ ) Ψ(r) ( Aˆ + B

= ΦA (r) + ΦB (r) ˆ Ψ(r) = Aˆ Ψ(r) + B

(198)

An example is given by the differential operator ∂ ∂2 + 3 + x4 2 ∂x ∂x

(199)

Another example is given by the kinetic energy operator for a particle in Cartesian coordinates  2  ¯h2 ∂ ∂2 ∂2 + + (200) Tˆ = − 2m ∂x2 ∂y 2 ∂z 2 where the kinetic energies for motion along the x, y and z axis add.

3.4.3

Operator Multiplication

ˆ each defined by their action on every Given two differential operators Aˆ and B, conceivable wave function Ψ(r) by Aˆ Ψ(r) ˆ ΦA (r) B

= ΦA (r) = ΦBA (r)

(201)

then the product of two operators is defined by their successive actions. ˆ Aˆ Ψ(r) B

ˆ ΦA (r) = B = ΦBA (r)

62

(202)

n

m

∂ ∂ Two examples are given by multiplying two operators ∂x n and ∂xm which n+m ∂ yields the compound operator ∂x n+m and similarly by multiplying the operan m n+m tors x and x yields x .

Important examples of the multiplication of two operators are given by : The kinetic energy for motion in one dimension. The kinetic energy operator is derived from the classical expression Tˆx

= = =

pˆ2x 2m pˆx . pˆx 2m    1 ∂ ∂ − ih ¯ − ih ¯ 2m ∂x ∂x

= −

¯h2 ∂ 2 2 m ∂x2

(203)

The kinetic energy for motion in three dimensions. The kinetic energy operator is derived as Tˆ

p ˆ2 2m p ˆ.p ˆ = 2m  2 1 ∂ ∂ ∂ = − i ¯h eˆx − i ¯h eˆy − i ¯h eˆz 2m ∂x ∂y ∂z   2 2 2 2 h ¯ ∂ ∂ ∂ = − + + 2m ∂x2 ∂y 2 ∂z 2

=

(204)

where we have used the representation of the vector momentum p in terms of its components and the unit vectors eˆx , eˆy and eˆz . ˆ is given by the vector The (vector) orbital angular momentum operator L product of two vector operators ˆ = rˆ ∧ p ˆ L

(205)

ˆ x + eˆy L ˆ y + eˆz L ˆz ˆ = eˆx L L

(206)

and is expressed as

63

where ˆz L ˆy L ˆx L



 ∂ ∂ = − i ¯h x − y ∂y ∂x   ∂ ∂ = −ih ¯ z − x ∂x ∂z   ∂ ∂ = − i ¯h y − z ∂z ∂y

(207)

This is actually a pseudo-vector, as it transforms like a vector under every symmetry operation except inversion. It should be noted that in general, unlike in the previous two examples, the product of two operators depends on the order in which they act on the wave function. If two operators have the effect ˆ Ψ(r) B Aˆ ΦB (r)

= ΦB (r) = ΦAB (r)

(208)

then the product of the two operators is defined by their successive actions ˆ Ψ(r) Aˆ B

= Aˆ ΦB (r) = ΦAB (r) 6= ΦBA (r)

ˆ = xm one has and B   ∂ m ˆ ˆ A B Ψ(x) = x Ψ(x) ∂x ∂ = m xm−1 Ψ(x) + xm Ψ(x) ∂x

For example, in multiplying Aˆ =

(209)

∂ ∂x

(210)

whereas on taking the product in the opposite order one obtains a different result   ˆ Aˆ Ψ(x) = xm ∂ B Ψ(x) (211) ∂x

3.4.4

Commutators

ˆ is defined as the difference of the The commutator of two operators Aˆ and B products of two operators taken in different orders ˆ ] = Aˆ B ˆ − B ˆ Aˆ [ Aˆ , B

64

(212)

In this equation, like other operator equations, it should always be remembered that the operators are assumed to act on an arbitrary wave function. Note that the commutator is anti-symmetric ˆ ] = −[B ˆ , Aˆ ] [ Aˆ , B

From the previous example, where Aˆ = ˆ ] Ψ(x) [ Aˆ , B

∂ ∂x

(213)

ˆ = xm , one has and B

ˆ Ψ(x) − B ˆ Aˆ Ψ(x) = Aˆ B     ∂ ∂ = xm Ψ(x) − xm Ψ(x) ∂x ∂x = m xm−1 Ψ(x)

(214)

Since the above equation is true for every arbitrary Ψ(x), one has the operator equation ∂ ∂ xm − xm = m xm−1 (215) ∂x ∂x where the presence of the wave function is implicitly assumed. An important ∂ example is given by the commutator of x ˆ = x and pˆx = − i ¯h ∂x then [x ˆ , pˆx ] = [ x , − i ¯h

∂ ] = i ¯h ∂x

(216)

The commutators of the x, y, and z components of the momentum and position operators are given by [x ˆ , pˆx ] = [ yˆ , pˆy ] = [ zˆ , pˆz ] = i ¯h

(217)

If the commutator of two operators is zero, then the two operators are said to commute. An example of two commuting operators is given by x ˆ = x and ∂ pˆy = − i ¯ h ∂y . In general, the commutators between different components of the particle’s coordinates and momenta are zero. [x ˆ , pˆy ] = [ yˆ , pˆx ] = 0 [ yˆ , pˆz ] = [ zˆ , pˆy ] = 0 [x ˆ , pˆz ] = [ zˆ , pˆx ] = 0

(218)

ˆ and Cˆ The commutator of an operator Aˆ and the sum of two operators B is given by ˆ + Cˆ ] = [ Aˆ , B ˆ ] + [ Aˆ , Cˆ ] [ Aˆ , B (219)

65

ˆ and whereas the commutator of an operator and a product of two operators B Cˆ is given by ˆ Cˆ ] = [ Aˆ , B ˆ ] Cˆ + B ˆ [ Aˆ , Cˆ ] [ Aˆ , B

(220)

ˆ , Cˆ ] = Aˆ [ B ˆ , Cˆ ] + [ Aˆ , Cˆ ] B ˆ [ Aˆ B

(221)

and

Repeated application of the above relations allows one to express the comˆ n ] as mutator of [ Aˆ , B ˆn ] [ Aˆ , B

ˆ ]B ˆ n−1 + B ˆ [ Aˆ , B ˆ ]B ˆ n−2 + B ˆ 2 [ Aˆ , B ˆ ]B ˆ n−3 = [ Aˆ , B ˆ n−1 [ Aˆ , B ˆ ]B ˆ + B ˆ n−1 [ Aˆ , B ˆ ] + ..... + B (222)

which involves the sum of n commutators. ˆ ] is not an operator but is a constant instead, If the commutator [ Aˆ , B the above expression simplifies to ˆ n ] = n [ Aˆ , B ˆ ]B ˆ n−1 [ Aˆ , B

(223)

ˆ ] is a constant then, This can be used to prove that if the commutator [ Aˆ , B ˆ for any function f (B) which has a power series expansion, one has ˆ ] = [ Aˆ , B ˆ ] f 0 (B) ˆ [ Aˆ , f (B)

(224)

where f 0 (x) is the derivative of f (x), that is f 0 (x) = ∂f ∂x . Examination of this operator equation gives justification for representing any pair of operators that have a constant commutator as a pair which consists of the variable and the commutation constant times a derivative w.r.t. the variable. ——————————————————————————————————

3.4.5

Exercise 13

Determine the commutator [ p ˆ , V (ˆ r) ]. This expression occurs in the equation which determines the time derivative of the average value of the momentum, p. ——————————————————————————————————

3.4.6

Solution 13

The commutators between the components of momentum and the components of the position are given by [ pˆi , xj ] = − i ¯h δi,j 66

(225)

These commutators are constants. Thus, on defining the potential as a Taylor series expansion in r one has ∂V (r) ∂xi

(226)

[p ˆ , V (r) ] = − i ¯h ∇ V (r)

(227)

[ pˆi , V (r) ] = − i ¯h which is the i-th component of the commutator

The commutator of the momentum and the potential is related to the gradient of the potential and thus is related to the force. ——————————————————————————————————

3.4.7

Exercise 14

ˆ z ] , [ yˆ , L ˆ z ] and [ zˆ , L ˆ z ] of the Determine the three commutators [ x ˆ, L components of the position vector and the z component of the angular momentum operator. ˆ z with the components of the momentum pˆ , Also find the commutator of L ˆ ˆ ˆz ] . [ pˆx , Lz ] , [ pˆy , Lz ] and [ pˆz , L ——————————————————————————————————

3.4.8

Solution 14

The z component of the angular momentum is given by ˆz = x L ˆ pˆy − yˆ pˆx

(228)

Then, the commutator with the x component of the position is given by ˆz ] = [ x [x ˆ, L ˆ, x ˆ pˆy ] − [ x ˆ , yˆ pˆx ] = − yˆ [ x ˆ , pˆx ] = −ih ¯ yˆ

(229)

The commutator with the y component of the position is given by ˆz ] [ yˆ , L

= [ yˆ , x ˆ pˆy ] − [ yˆ , yˆ pˆx ] = +x ˆ [ yˆ , pˆy ] = + i ¯h x ˆ

(230)

The third commutator is given by ˆ z ] = [ zˆ , x [ zˆ , L ˆ pˆy ] − [ x ˆ , yˆ pˆx ] = 0 67

(231)

Hence, we obtain [x ˆ, [ yˆ , [ zˆ ,

ˆ z ] = − i ¯h yˆ L ˆ z ] = + i ¯h x L ˆ ˆ Lz ] = 0

(232)

The remaining commutators can be obtained from the two non-zero commutators by cyclically permuting x, y and z. Similarly, the commutation relations between the components of the angular momentum and the components of the momentum are evaluated from ˆz ] [ pˆx , L

= [ pˆx , x ˆ pˆy ] − [ pˆx , yˆ pˆx ] = [ pˆx , x ˆ ] pˆy = − i ¯h pˆy

(233)

Hence, the commutators of the z component of angular momentum with the components of the momentum are found as ˆ z ] = − i ¯h pˆy [ pˆx , L ˆ z ] = + i ¯h pˆx [ pˆy , L ˆz ] = 0 [ pˆz , L

(234)

The other commutators can be found by cyclic permutation of the subscripts x, y and z. ——————————————————————————————————

3.4.9

Exercise 15

ˆ ] ], where the Hamiltonian is Determine the double commutator [ x , [ x , H pˆ2 ˆ = given by H 2 m + V (r). This commutator forms the basis of the ThomasReiche-Kuhn Sum rule for the intensity of optical absorption by atoms. ——————————————————————————————————

3.4.10

Solution 15

The commutator between the x component of the position and the Hamiltonian ˆ can be decomposed into the sum of the commutators between x and the H pˆ2 potential V (r) and the commutator of x with the potential energy 2m 2 ˆ ] = [ x , V (r) ] + [ x , pˆ ] [x, H 2m

68

(235)

The commutator between the x component of the position and the potential vanishes [ x , V (r) ] = 0 (236) The only non-zero part of the commutator is the commutator between the x component of the kinetic energy and the position. This non-zero commutator is evaluated as   pˆ2x 1 [x, ] = [ x , pˆx ] pˆx + pˆx [ x , pˆx ] 2m 2m i ¯h = 2 pˆx (237) 2m Hence, the double commutator is given by 2 ˆ ] ] = − ¯h [x, [x, H m

(238)

—————————————————————————————————— Another special example of an operator equation that we shall extensively use is in the definition of an exponential of an operator   X λn Aˆn ˆ (239) exp λ A = n! n where λ is any complex number. Consider the operator which is a function of λ and defined via, fˆ(λ) = eλ

ˆ A

ˆ e− B

ˆ λ A

(240)

then it is easy to prove that the derivatives are given by the series of expressions,  ˆ df = Aˆ fˆ(λ) − fˆ(λ) Aˆ dλ = [ Aˆ , fˆ(λ) ]  2 ˆ d f dfˆ = [ Aˆ , ] 2 dλ dλ = [ Aˆ , [ Aˆ , fˆ(λ) ] ] (241) This infinite set of equations can be used in the Taylor expansion of fˆ(λ) to yield eλ

ˆ A

ˆ e− B

ˆ λ A

ˆ + λ [ Aˆ , B ˆ ] + = B +

69

λ2 ˆ ˆ ]] [ A , [ Aˆ , B 2! λ3 ˆ ˆ ] ] ] + ... [ A , [ Aˆ , [ Aˆ , B 3! (242)

ˆ where we have used the fact that fˆ(0) = B. ——————————————————————————————————

3.4.11

Exercise 16

Given the first order differential equation i ¯h

ˆ (t) dU ˆ U ˆ (t) = H dt

(243)

ˆ (0) = 1, find an expression for U ˆ (t). This equation and operator is and U usually seen in the context of the time evolution of a wave function, and the initial condition expresses the fact that the wave function is continuous at t = 0. ——————————————————————————————————

3.4.12

Solution 16

The solution for the time evolution operator is found by integrating the equation Z t ˆ U ˆ (t0 ) ˆ (t) = U ˆ (0) − i dt0 H (244) U ¯h 0 and then iterating ˆ (t) U

ˆ (0) − i = U ¯h

Z

t

0

Z t ˆ (0) − i = U ¯h 0 ˆ (0) − i t H ˆ = U ¯h  it ˆ = exp − H ¯h

ˆ U ˆ (0) − 1 dt H ¯h2 0

Z

t

dt 0

0

Z

t0

ˆ H ˆ U ˆ (t”) dt” H

0

Z t Z t0 0 ˆ H ˆ U ˆ (0) + . . . ˆ U ˆ (0) − 1 dt0 H dt dt” H ¯h2 0 0 2 ˆ (0) − 1 t H ˆ H ˆ U ˆ (0) + . . . U 2 2 h ¯  ˆ (0) U (245)

In the last line the series is recognized as the expansion of the exponential. Since ˆ (0) = 1, we obtain the solution as the initial condition corresponds to U   ˆ (t) = exp − i t H ˆ U (246) ¯h

——————————————————————————————————

70

3.4.13

Exercise 17

Given the first order linear differential operator equation ˆ dB(λ) ˆ = [ Aˆ , B(λ) ] dλ ˆ ˆ find an expression for B(λ). ˆ with the boundary condition B(0) = C,

(247)

——————————————————————————————————

3.4.14

Solution 17

Since

ˆ ∂ B(λ) ˆ = [ Aˆ , B(λ) ] ∂λ one can integrate the equation to yield Z λ ˆ ˆ ˆ 0) ] B(λ) − B(0) = dλ0 [ Aˆ , B(λ

(248)

0

ˆ B(λ) − Cˆ =

Z

λ

ˆ 0) ] dλ0 [ Aˆ , B(λ

(249)

0

This equation can be iterated to yield Z λ Z ˆ B(λ) = Cˆ + dλ0 [ Aˆ , Cˆ ] + 0

ˆ B(λ)

= Cˆ +

Z

λ

dλ0

0 λ

dλ [ Aˆ , Cˆ ] + 0

0

Z

λ0

ˆ dλ” [ Aˆ , [ Aˆ , B(λ”) ]]

0 λ

dλ 0

Z

0

Z

λ0

dλ” [ Aˆ , [ Aˆ , Cˆ ] ] + . . .

0

(250) which can be evaluated as λ2 ˆ ˆ B(λ) = Cˆ + λ [ Aˆ , Cˆ ] + [ A , [ Aˆ , Cˆ ] ] + . . . 2! ˆ which is the Taylor series of B(λ).

(251)

From eqn(242) of the previous example, the Taylor series can be recognized as the expansion of     ˆ ˆ ˆ ˆ B(λ) = exp + λ A C exp − λ A (252) Alternatively, the series can be re-arranged as     λ2 ˆ2 λ2 ˆ2 ˆ ˆ ˆ ˆ B(λ) = 1 + λA + A + ... C 1 − λA + A + ... 2! 2!     = exp + λ Aˆ Cˆ exp − λ Aˆ (253)

71

giving the same result. ——————————————————————————————————

3.4.15

Exercise 18

The Baker - Campbell - Hausdorf relation6         ˆ ˆ ) exp 1 [ Aˆ , B ˆ ] exp Aˆ exp B = exp ( Aˆ + B 2

(254)

ˆ that have commutator which is a constant, is only valid for operators Aˆ and B i.e. ˆ ] = Const. [ Aˆ , B (255) Derive the Baker - Cambell - Hausdorff relation. ——————————————————————————————————

3.4.16

Solution 18

First we note that, from the series expansion of the exponential, one can show       ˆ ˆ ˆ ˆ , Aˆ ] exp λ B , A = λ exp λ B [ B (256) Now let us consider the operator function fˆ(λ) defined by       ˆ exp − λ ( Aˆ + B ˆ ) fˆ(λ) = exp λ Aˆ exp λ B

(257)

Then the derivative with respect to λ is given by ∂ fˆ ∂λ



    ˆ exp − λ ( Aˆ + B ˆ ) Aˆ exp λ B       ˆ B ˆ exp − λ ( Aˆ + B ˆ ) + exp λ Aˆ exp λ B       ˆ ( Aˆ + B ˆ ) exp − λ ( Aˆ + B ˆ ) − exp λ Aˆ exp λ B

=

exp

λ Aˆ



(258) 6 J.E.

Campbell, Proc. London Math. Soc. 29, 14 (1898). H.F. Baker, London Math. Soc. Ser. 3, 24 (1904). F. Hausdorf, Ber. Vehr. Sachs. Wissen. Leipzig, Math.-Naturwiss. Kl. 58, 19 (1906).

72

ˆ ] with Aˆ in the first term, we find On commuting exp[ λ B ∂ fˆ = λ exp ∂λ



λ Aˆ



 exp

ˆ λB



ˆ ] exp [ Aˆ , B



ˆ ) − λ ( Aˆ + B

 (259)

as the other terms cancel identically. Furthermore as the commutator is a constant it may be factored out, thus we have ∂ fˆ ∂λ



ˆ ] = λ [ Aˆ , B

(260)

which on integrating and re-arranging yields        2  ˆ ˆ ) exp λ [ Aˆ , B ˆ] exp λ Aˆ exp λ B = exp λ ( Aˆ + B (261) 2 and on setting λ = 1 we have proved the relationship. —————————————————————————————————— The Taylor-MacLaurin expansion of a function f (x) is given by   ∂ f (x + a) = exp a f (x) ∂x which has the three-dimensional analogue   f (r + a) = exp a . ∇ f (r)

(262)

(263)

This is equivalent to a shift of the origin by + a, or a shift of the function by − a.

3.4.17

Eigenvalue Equations

Given a differential operator Aˆ one can find eigenfunctions ϕa (r) and eigenvalues a. The eigenfunctions and eigenvalues are determined from the eigenvalue equation, (264) Aˆ ϕa (r) = a ϕa (r)

An example is given by the eigenfunctions of the x component of the momentum, pˆx , defined on the interval (−∞,∞). The eigenvalue equation is pˆx ϕp (r) = px ϕp (r) ∂ −ih ¯ ϕp (r) = px ϕp (r) ∂x 73

(265)

The eigenfunctions are of the form  ϕp (r) ∝ exp

i

px x ¯h

 (266)

and the eigenvalue px is any real number. Thus, the eigenvalues form a continuous spectra with values in the interval (−∞,∞). The momentum eigenfunctions pˆ2 are also eigenfunctions of the kinetic energy operator Tˆx , since Tˆx = 2 xm . We should note that two eigenfunctions correspond to the same eigenvalue of Tˆx . These two eigenfunctions are those which correspond to the momentum eigenvalues + px and − px . When an eigenvalue corresponds to more than one eigenfunction, the eigenvalue is said to be degenerate. The above eigenfunctions of Tx are doubly degenerate. The degeneracy is the number of linearly independent7 eigenfunctions corresponding to the same eigenvalue, and this depends on the space of the functions being considered. For example, if the wave functions depend on x, y and z, one can find eigenfunctions of pˆx of the form  Ψ(x, y, z) =

1 2 π ¯h

 32



px x exp i ¯h





py y exp i ¯h





pz z exp i ¯h

 (267)

the eigenvalue px is infinitely degenerate. d

Note that a factor of h ¯ − 2 has been introduced into the normalization of the d-dimensional momentum eigenstates, or equivalently the d-dimensional Fourier transform. This factor was introduced in order that the momentum distribution function is properly normalized. In general, given an eigenfunction Ψa (r) of an operator Aˆ with eigenvalue ˆ satisfies the eigenvalue equation a, then any function of the operator f (A) ˆ Ψa (r) = f (a) Ψa (r) f (A)

(268)

where the eigenvalue is f (a). Another example is given by the energy eigenfunctions of the one-dimensional ˆ harmonic oscillator, which has the Hamiltonian operator H ¯ 2 ∂2 m ω2 2 ˆ = − h H + x 2 2 m ∂x 2

(269)

7 A set of functions φ is said to be linearly independent if any function of the set cannot n be expressed as a linear superposition of the other members of the set. That is, if the set of φn is linearly independent, the only solution of the equation

X

Cn φn = 0

n

is the trivial solution in which all the coefficients are zero Cn = 0.

74

The energy eigenvalue equation for the harmonic oscillator is   ¯h2 ∂ 2 m ω2 2 ˆ Ψn (x) = H − + x Ψn (x) 2 m ∂x2 2 = En Ψn (x)

(270)

which has the energy eigenvalues En = ¯h ω ( n + 12 ) where n is any positive integer ( including 0 ). The energy eigenfunctions are Ψn (x), where ( − 1 )n Ψn (x) = 1 2n n! 2



mω hπ ¯

 14

mω 2 exp[ + x ] 2 ¯h



¯ h mω

 n2

mω 2 dn exp[ − x ] n dx ¯h

(271) In this case the eigenvalues are in the form of a discrete set of numbers, and the eigenfunctions are localized about the origin. The harmonic oscillator, classically, corresponds to a situation in which a particle would be bound to the origin by the quadratic potential, for any value of the energy E. ——————————————————————————————————

3.4.18

Exercise 19

Show that if

 Ψ(r) = exp

 − φ(r)

(272)

satisfies appropriate boundary conditions then it represents a bound state wave function with energy eigenvalue E for a particle moving in the potential   h2 ¯ V (r) = ( ∇φ(r) )2 − ∇2 φ(r) + E (273) 2m

——————————————————————————————————

3.4.19

Exercise 20

An electron is moving in one dimension and is constrained to the region outside a perfect conductor x > 0, where the potential is given by the image potential V (x) = −

e2 2x

(274)

where e is the charge of the electron. Find a bound state energy eigenfunction φ(x) and the average value of x.

75

Image potential for electrons near a metal surface

V(x) x0 / e

2

1 0.5 0 -0.5 2

-1

V(x) = - e / 2 x

-1.5 -2 -2

0

2

4

6

8

10

x / x0

Figure 17: The one-dimensional image potential V (x) which confines an electron close to the surface of a metal. The bound state energy is marked by a horizontal line. Hint: Try a wave function of the form x exp[ − α x ] for x > 0. ——————————————————————————————————

3.4.20

Solution 20

The trial wave function is chosen as  φ(x) = A x exp

 − αx

(275)

for x > 0 and is zero for negative values of x, and is shown in fig(18). Thus, the wave function is localized as x → ∞ and vanishes at the surface at x = 0, representing a particle trapped in the attractive image potential. The electron does not penetrate into the metal. On substituting the wave function into the energy eigenvalue equation   e2 h2 ∂ 2 ¯ − − φ(x) = E φ(x) (276) 2 m ∂x2 2x and on noting that ∂φ 1 = ( − α ) φ(x) ∂x x 76

(277)

Bound state wave function 0.6

Ψ0(x)

0.5 0.4 0.3 0.2 0.1 0 -2

0

2

4

x / x0

6

8

10

Figure 18: The wave function Ψ(x) of an electron bound to the surface of a metal. and the second derivative is given by ∂2φ 2 = −( − α ) α φ(x) ∂x2 x one finds that the eigenvalue equation reduces to   h2 ¯ α e2 2 − (α − 2 ) − φ(x) = E φ(x) 2m x 2x

(278)

(279)

This is an algebraic equation, which is solved by choosing α such that m e2 2 ¯h2 and the energy eigenvalue E is then given by α =

(280)

m e4 (281) 8 ¯h2 The eigenvalue E is negative, corresponding to the fact that classically the electron is trapped in the potential well and does not have enough energy to reach x → ∞. The threshold energy for the particle to reach infinity is E = 0, and hence the particle must have a negative energy. E = −

——————————————————————————————————

77

3.4.21

Exercise 21

A particle of mass m is confined to the region x > 0 and is subjected to a potential V (x) given by  2  a a V (x) = V0 − (282) x2 x where V0 and a are constants. The potential is sketched in fig(19). Derive the

1

2

V(x) = V0 [ (a/x) - (a/x) ]

V(x)

0.5

0 0

2

4

6

8

10

-0.5

x/a

Figure 19: The potential V (x) for a one-dimensional hydrogen-like atom. bound state energy and wave function. Hint: Try a wave function of the form xs exp[ − α x ] for x > 0. ——————————————————————————————————

3.4.22

Solution 21

We shall consider the trial wave function   φ(x) = C xs exp − α x

78

(283)

which has the second derivative given by   ∂2 s(s − 1) s 2 φ(x) = − 2α + α φ(x) ∂x2 x2 x

(284)

On substituting into the energy eigenvalue equation one finds the consistency equations 2 m V0 2 a ¯h2 2 m V0 a = ¯h2 2mE = − ¯h2

s(s − 1)

=

2sα α2

(285)

Thus, we find the exponent s as 1 s = ± 2

r

2 m V0 2 1 + a 4 ¯h2

(286)

Since the wave function should be square integrable we take the positive root. Then from the second equation we find 2 m V0 h ¯2

α = 1 +

q

1 +

a 8 m V0 h ¯2

(287) a2

from which one can find the energy E. ——————————————————————————————————

3.4.23

Exercise 22

Prove that the energy eigenfunctions of the harmonic oscillator are given by Ψn (x) given above. ——————————————————————————————————

3.4.24

Solution 22

The wave function Ψn (x) should satisfy the energy eigenvalue equation,   h2 ∂ 2 ¯ m ω2 2 1 − + x Ψn (x) = h ¯ω ( n + ) Ψn (x) (288) 2 2 m ∂x 2 2 First let us transform to the dimensionless variable y where r mω y = x ¯h 79

(289)

Then, the energy eigenvalue equation can be simplified to   ∂2 2 − + y Φn (y) = ( 2 n + 1 ) Φn (y) ∂y 2 We have to show that Φn (y) given by    n   y2 ∂ Φn (y) = exp + exp − y 2 2 ∂y

(290)

(291)

satisfies the above dimensionless eigenvalue equation. We shall prove this by induction. We first note that the above trial wave function with n = 0 satisfies the eigenvalue equation as       ∂2 y2 y2 2 − + y exp − = exp − (292) ∂y 2 2 2 So the trial wave function satisfies the equation for n = 0, and is thus a solution for this particular n. Then we shall prove that, if the trial wave function satisfies the eigenvalue equation for an arbitrary n, then the function also satisfies the equation for ( n + 1 ). Let us first note that    n   y2 ∂ y2 exp − exp + 2 ∂y 2        n 2 ∂ y2 y = exp + exp − 2 ∂y 2  n ∂ = + − y ∂y The trial wave function is assumed   ∂2 2 − + y + ∂y 2  = (2n + 1) +

to satisfy the equation n  ∂ − y exp − ∂y n  ∂ − y exp − ∂y

 y2 2  y2 2

(293)

(294)

and we aim to show that if the above equation is true then the equation with n + 1   n+1   ∂2 ∂ y2 2 − + y + − y exp − ∂y 2 ∂y 2  n+1   ∂ y2 = (2n + 3) + − y exp − (295) ∂y 2 80

is also true. We note that the operator identity       ∂2 ∂ ∂ ∂2 2 2 − + y + − y = + − y − + y + 2 ∂y 2 ∂y ∂y ∂y 2 (296) is always valid. On substituting this identity into the eigenvalue equation with the n + 1 th trial function, we have    n   ∂ ∂2 ∂ y2 2 + − y − + y + 2 + − y exp − ∂y ∂y 2 ∂y 2   n    ∂ ∂ y2 − y (2n + 3) + − y exp − = + ∂y ∂y 2 (297) and on cancelling the term    n   ∂ ∂ y2 + − y 2 + − y exp − ∂y ∂y 2 from both sides of the equation, we obtain    ∂ ∂2 2 + − y − + y + ∂y ∂y 2    ∂ = + − y (2n + 1) + ∂y

(298)

n   ∂ y2 − y exp − ∂y 2 n   ∂ y2 − y exp − ∂y 2 (299)

This is true if eqn(294) holds true, as we are only taking a derivative of eqn(294). Equation(294) is true for n = 0 and so from the above we have shown that it is true for n = 1. By induction this equation is also true for all higher integer values of n. —————————————————————————————————— ˆ which corresponds to a physical measurement A and Given an operator, A, a system that has a wave function that is an eigenfunction of the operator corresponding to the eigenvalue a then a measurement of that physical quantity A definitely results in the value a.

3.4.25

Adjoint or Hermitean Conjugate Operators

Given an operator Aˆ one can construct its adjoint or Hermitean conjugate, Aˆ† . The adjoint operator or Hermitean conjugate operator is defined by the matrix elements between two arbitrary wave functions  Z ∗ Z ∗ 3 ∗ † 3 ˆ ˆ d r Φ (r) A Ψ(r) = d r Ψ (r) A Φ(r) (300)

81

This definition involves the inner product of the wave function Aˆ Ψ(r), which ˆ and Φ(r). The defiis the state Ψ(r) after it is transformed by the operator A, nition relates this inner product to the complex conjugate of the inner product of the state Aˆ† Φ(r) with Ψ(r). From the definition of the adjoint or Hermitean conjugate, it is easy to prove that the Hermitean conjugate of the Hermitean conjugate of an operator Aˆ is ˆ the same as the original operator A, 

Aˆ†

†

= Aˆ

(301)

This is proved by using the definition of the adjoint twice  Z ∗ Z d3 r Ψ∗ (r) Aˆ† Φ(r) d3 r Φ∗ (r) Aˆ Ψ(r) =  Z

3



d r Φ (r)

=



Aˆ†

†

 ∗ ∗ Ψ(r) (302)

which, since (z ∗ )∗ ˆ (Aˆ† )† = A.

=

z for any complex number, yields the identification

Also a complex constant a, when regarded as an operator, has a Hermitean conjugate which is just its complex conjugate a∗ . An example of finding the Hermitean conjugate of an operator is given by considering the differential operator in one dimension. Aˆ =

∂ ∂x

(303)

The Hermitean conjugate or self adjoint operator, Aˆ† , is found from Z +∞ ∂ dx Φ∗ (x) Ψ(x) = ∂x −∞ ∞   Z +∞ ∂ = Φ∗ (x) Ψ(x) − dx Φ∗ (x) Ψ(x) ∂x −∞ −∞   Z +∞ ∂ = − dx Φ∗ (x) Ψ(x) (304) ∂x −∞ where the second line involves integration by parts and the third line utilizes the boundary conditions on the wave functions lim

|x| → ∞

Ψ(x) = 0

82

(305)

which are implied by the normalization condition. From this, one identifies the Hermitean conjugate operator as ∂ Aˆ† = − ∂x

(306)

Another example of a Hermitean conjugate of an operator is given by translation operator, which translates the wave function by the vector a.   ˆ S(a) = exp − a . ∇ (307) This, when acting on a wave function Ψ(r) generates the Taylor MacLaurin series expansion of the function Ψ(r − a). The Hermitean conjugate is found from the series expansion of the exponential operator, and integrating by parts to yield   Sˆ† (a) = exp + a . ∇ (308) ˆ which coincides with the inverse of S(a), as ˆ ˆ Sˆ† (a) S(a) = S(a) Sˆ† (a) = 1

(309)

Sˆ† (a) = Sˆ−1 (a)

(310)

Thus, we have Operators that have their Hermitean conjugate operators equal to their inverses are called unitary operators. Thus, an operator Aˆ is a unitary operator if Aˆ† = Aˆ−1

(311)

ˆ Hence, S(a) is a unitary operator. ˆ (t) Another unitary operator is given by the time translation operator U which is found to be  ˆ t  H ˆ U (t) = exp − i (312) ¯h ˆ is the Hamiltonian operator where H ¯ 2 ∂2 ˆ = − h H + V (x) 2m ∂x2

(313)

and V (x) is a real function. Then it can be shown that the Hermitean conjugate ˆ is given by of U  ˆ  ˆ † (t) = exp + i H t U (314) ¯h

83

Translation through a distance a 1

φ(x)

φ(x-a)

φ'(x)=φ(T-1x)=φ(x-a)

0.8 0.6 0.4 0.2 0

x'=Tx=x+a x0 0

1

x0+a 2

3

0

1

2

3

x

x

ˆ Figure 20: The unitary operator S(a) produces a translation of the wave function by a. The translation operator Tˆ acts on the coordinates and transforms the point x to the point x0 = x + a, i.e. x0 = Tˆ x. Under the translation, the wave function φ(x) is transformed to φ0 (x), such that the value of φ0 at x0 is the same as the value of φ at the point x. Therefore, φ0 (x0 ) = φ(x) so, φ0 (x) = φ(Tˆ−1 x). The unitary operator which transforms the wave function ˆ is then specified as φ0 (x) = S(a) φ(x) = φ(Tˆ−1 x). In this case, the Hermitean conjugate is just the complex conjugate of the operator. Also we have that the Hermitean conjugate coincides with the inverse operator ˆ † (t) U ˆ (t) = U ˆ (t) U ˆ † (t) = 1 U (315) This is the condition that has to be satisfied for an operator to be a unitary operator. The unitary condition for the time evolution operator ensures that the normalization of the states is independent of time. ˆ )† is given The Hermitean conjugate of a product of two operators ( Aˆ B ˆ † Aˆ† ). by the product of the Hermitean conjugates taken in reverse order, ( B This can be proved by starting with the definition of the Hermitean conjugate of the product   ∗  † Z Z ˆ Φ(r) ˆ d3 r Ψ∗ (r) Aˆ B = d3 r Φ∗ (r) Aˆ B Ψ(r) (316) The product operator is defined in terms of the state which results when the 84

ˆ Φ(r). On using the definition of the Hermitean operator Aˆ acts on the state B conjugate twice, successively, one finds Z   ∗ 3 ∗ ˆ ˆ = d r Ψ (r) A B Φ(r)  ∗   Z ˆ Φ(r) = d3 r B Aˆ† Ψ(r)  ∗   ∗ Z ˆ Φ(r) = d3 r Aˆ† Ψ(r) B   Z ˆ † Aˆ† Ψ(r) = d3 r Φ∗ (r) B Z ˆ † Aˆ† Ψ(r) = d3 r Φ∗ (r) B (317) which completes the proof.

3.4.26

Hermitean Operators

A Hermitean operator (or self Adjoint operator) is any operator that satisfies the condition Aˆ† = Aˆ (318) That is, if an operator is Hermitean, the Hermitean conjugate of the operator is equal to the operator. Examples of Hermitean operators are given ∂ mentum operators pˆx = − i ¯h ∂x , pˆy = − i also the single-particle Hamiltonian defined by  2 h2 ¯ ∂ ∂2 ˆ H = − + + 2m ∂x2 ∂y 2

by, the components of the mo∂ ∂ , and ¯ ∂y h and pˆz = − i ¯h ∂z ∂2 ∂z 2

 + V (r)

(319)

is Hermitean. In spherical polar coordinates (r,θ,ϕ), the canonical generalized momenta could correspond to the Hermitean operators pˆr pˆθ pˆϕ

ih ¯ ∂ r r ∂r 1 i ¯h ∂ = − sin 2 θ 1 ∂θ 2 sin θ ∂ = − i ¯h ∂ϕ = −

(320)

These operators are Hermitean operators. Note that when these operators are taken between two wave functions, and integrated over the volume r2 dr sin θ dθ dϕ 85

the resulting factors of r and sin θ in the expressions are symmetrically placed with respect to the differential operator. ——————————————————————————————————

3.4.27

Exercise 23

Prove that the canonical momenta in spherical polar coordinates given in eqn(320) are Hermitean operators. ——————————————————————————————————

3.4.28

Exercise 24

The expression for the kinetic energy in the Hamiltonian of eqn(39) when quantized involves the sum of the three operators pˆ2r pˆ2θ pˆ2ϕ

¯2 ∂ 2 ∂ h r r2 ∂r ∂r ∂ ¯h2 ∂ sin θ = − sin θ ∂θ ∂θ 2 2 ∂ = − ¯h ∂ϕ2

= −

(321)

Show that these three operators are Hermitean. ——————————————————————————————————

3.4.29

Solution 24

The Hermitean conjugate of pˆ2ϕ is calculated from the matrix elements Z

2 π

Z dϕ

= −¯ h

2

Z

0 2 π

Z



dθ sin θ

Z dϕ

0

π

0 π

Z

dr r2 Φ∗ (r, θ, ϕ)

dθ sin θ 0

dr r2 Φ∗ (r, θ, ϕ) pˆ2ϕ Ψ(r, θ, ϕ)

0 ∞

0

∂2 Ψ(r, θ, ϕ) ∂ϕ2 (322)

by integrating by parts twice with respect to ϕ. On integrating by parts for the first time, one obtains Z 2π Z π Z ∞ ∂2 2 − ¯ h dϕ dθ sin θ dr r2 Φ∗ (r, θ, ϕ) Ψ(r, θ, ϕ) ∂ϕ2 0 0 0 86

= −¯ h

2

Z

π

dθ sin θ 0

+ ¯h2

Z

2 π

Z dϕ

0

 2 π ∂ dr r Φ (r, θ, ϕ) Ψ(r, θ, ϕ) ∂ϕ 0 0     Z ∞ ∂ ∗ ∂ Φ (r, θ, ϕ) Ψ(r, θ, ϕ) dθ sin θ dr r2 ∂ϕ ∂ϕ 0 (323)

Z

0

π



2





The boundary term vanishes since the wave functions are 2π periodic in ϕ, i.e., Ψ(r, θ, ϕ) = Ψ(r, θ, ϕ + 2π). On integrating by parts once again, one obtains Z 2π Z π Z ∞ ∂2 ∗ − ¯h2 dϕ dθ sin θ dr r2 Φ (r, θ, ϕ) Ψ(r, θ, ϕ) ∂ϕ2 0 0 0  Z 2π ∗ Z π Z ∞ ∂2 2 2 ∗ = −¯ h dϕ dθ sin θ dr r Ψ (r, θ, ϕ) Φ(r, θ, ϕ) ∂ϕ2 0 0 0 (324) where the boundary terms have vanished once again. Thus, we have found that  † ∂2 2 pˆϕ = − ¯h2 (325) ∂ϕ2 Hence, pˆ2ϕ is Hermitean. The Hermitean conjugate of pˆ2θ is calculated from the matrix elements Z 2π Z π Z ∞ dϕ dθ sin θ dr r2 Φ∗ (r, θ, ϕ) pˆ2θ Ψ(r, θ, ϕ) 0 0 0   Z 2π Z π Z ∞ ∂ ∂ 2 2 ∗ = −¯ h dϕ dθ dr r Φ (r, θ, ϕ) sin θ Ψ(r, θ, ϕ) ∂θ ∂θ 0 0 0 (326) by integrating by parts twice with respect to θ. Note that the factor sin θ in the metric for the volume integral has cancelled with a factor from the component of the kinetic energy. On integrating by parts for the first time, one obtains   Z 2π Z π Z ∞ ∂ ∂ 2 2 ∗ − ¯ h dϕ dθ dr r Φ (r, θ, ϕ) sin θ Ψ(r, θ, ϕ) ∂θ ∂θ 0 0 0   π Z 2π Z ∞ ∂ = −¯ h2 dϕ dr r2 Φ∗ (r, θ, ϕ) sin θ Ψ(r, θ, ϕ) ∂θ 0 0 0     Z 2π Z π Z ∞ ∂ ∗ ∂ 2 2 + ¯h dϕ dθ sin θ dr r Φ (r, θ, ϕ) Ψ(r, θ, ϕ) ∂θ ∂θ 0 0 0 (327) The boundary term vanishes since the wave functions are finite at the poles where sin θ = 0 . On integrating by parts once again, one obtains   Z 2π Z π Z ∞ ∂ ∗ 2 2 ∂ − ¯h dϕ dθ dr r sin θ Φ (r, θ, ϕ) Ψ(r, θ, ϕ) ∂θ ∂θ 0 0 0 87

= −¯ h

2

 Z

2 π

Z

π



Z





0

0

0

∂ dr r Ψ (r, θ, ϕ) ∂θ ∗

2



∂ sin θ Φ(r, θ, ϕ) ∂θ

 ∗ (328)

where the boundary terms have vanished once again. Thus, we have found that 

pˆ2ϕ

†

= −h ¯2

1 ∂ sin θ ∂θ

 sin θ

∂ ∂θ

 (329)

Hence, pˆ2ϕ is Hermitean. Note that the weight factor sin θ in the integral over the solid angle played a crucial role in the proof of Hermiticity. The radial component of the kinetic energy have the matrix elements Z 2π Z π Z ∞ dϕ dθ sin θ dr r2 Φ∗ (r, θ, ϕ) pˆ2r Ψ(r, θ, ϕ) 0 0 0   Z 2π Z π Z ∞ ∂ 2 ∗ 2 ∂ dθ sin θ dr Φ (r, θ, ϕ) = −¯ h dϕ r Ψ(r, θ, ϕ) ∂r ∂r 0 0 0 (330) by integrating by parts twice with respect to r. On integrating by parts for the first time, one obtains   Z 2π Z π Z ∞ ∂ ∂ r2 Ψ(r, θ, ϕ) − ¯h2 dϕ dθ sin θ dr Φ∗ (r, θ, ϕ) ∂r ∂r 0 0 0  ∞  Z 2π Z π ∂ = −¯ h2 dϕ dθ sin θ r2 Φ∗ (r, θ, ϕ) Ψ(r, θ, ϕ) ∂r 0 0 0     Z 2π Z π Z ∞ ∂ ∂ 2 2 ∗ Φ (r, θ, ϕ) Ψ(r, θ, ϕ) + ¯h dϕ dθ sin θ dr r ∂r ∂r 0 0 0 (331) The boundary term vanishes since the wave function vanishes as r → ∞, i.e., limr → ∞ Ψ(r, θ, ϕ) = 0 and the weight function vanishes at r = 0 . On integrating by parts once again, one obtains   Z 2π Z π Z ∞ ∂ ∂ ∗ − ¯ h2 dϕ dθ sin θ dr r2 Φ (r, θ, ϕ) Ψ(r, θ, ϕ) ∂r ∂r 0 0 0  Z 2π   ∗ Z π Z ∞ ∂ 2 ∗ 2 ∂ = −¯ h dϕ dθ sin θ dr Ψ (r, θ, ϕ) r Φ(r, θ, ϕ) ∂r ∂r 0 0 0 (332) where the boundary terms have vanished once again. Thus, we have found that 

pˆ2r

†

¯h2 ∂ = − 2 r ∂r 88



∂ r ∂r 2

 (333)

is Hermitean. In the proof, the weight factor r2 played a crucial role in showing the radial part of the kinetic energy is Hermitean. ——————————————————————————————————

3.4.30

Exercise 25

Using the expression for the momentum operator p = − i ¯h ∇ in spherical polar coordinates   ∂ 1 ∂ 1 ∂ p = −i¯ h eˆr + eˆθ + eˆϕ (334) ∂r r ∂θ r sin θ ∂ϕ where the unit vectors are given in terms of the Cartesian unit vectors via eˆr eˆθ eˆϕ

= sin θ cos ϕ eˆx + sin θ sin ϕ eˆy + cos θ eˆz = cos θ cos ϕ eˆx + cos θ sin ϕ eˆy − sin θ eˆz = − sin ϕ eˆx + cos ϕ eˆy

(335)

show that the resulting expression for the kinetic energy agrees with the expression inferred from the previous exercise. ——————————————————————————————————

3.4.31

Solution 25

First we shall note that the unit vectors are independent of r and are only functions of the angles, so ∂ˆ er ∂θ ∂ˆ er ∂ϕ ∂ˆ eθ ∂θ ∂ˆ eθ ∂ϕ ∂ˆ eϕ ∂θ ∂ˆ eϕ ∂ϕ

= eˆθ =

sin θ eˆϕ

= − eˆr =

cos θ eˆϕ

=

0 

= −

 sin θ eˆr + cos θ eˆθ

Then, we have 2

pˆ = − ¯ h

2

 eˆr

∂ ∂ ∂ 1 ∂ . eˆr + eˆr . eˆθ ∂r ∂r ∂r r ∂θ 89

(336)

∂ 1 ∂ 1 ∂ ∂ . eˆϕ + eˆθ . eˆr ∂r r sin θ ∂ϕ r ∂θ ∂r ∂ 1 ∂ 1 ∂ 1 ∂ 1 + eˆθ . eˆθ + eˆθ . eˆϕ r ∂θ r ∂θ r ∂θ r sin θ ∂ϕ ∂ ∂ 1 ∂ 1 1 ∂ . eˆr + eˆϕ . eˆθ + eˆϕ r sin θ ∂ϕ ∂r r sin θ ∂ϕ r ∂θ  ∂ ∂ 1 1 + eˆϕ . eˆϕ r sin θ ∂ϕ r sin θ ∂ϕ + eˆr

(337)

which on evaluating the derivatives and evaluating the scalar products of the orthogonal unit vectors becomes  2 ∂ ∂2 2 2 + pˆ = − ¯ h ∂r2 r ∂r 2 cos θ ∂ 1 ∂ + + 2 r2 ∂θ2 r sin θ ∂θ 1 ∂2 + (338) 2 2 r sin θ ∂ϕ2 The terms containing the first derivatives come from the change in the unit vectors of the fourth, seventh and eight terms of the previous equation. Then this can be simplified to yield       1 ∂ 1 ∂ ∂ 1 ∂2 2 2 2 ∂ pˆ = − ¯ h r + 2 sin θ + 2 2 r2 ∂r ∂r r sin θ ∂θ ∂θ r sin θ ∂ϕ2 (339) which involves the appropriate expression for the Laplacian in spherical polar coordinates. —————————————————————————————————— ˆ which is not Hermitean, i.e. B ˆ = ˆ † , then Given an arbitrary operator B 6 B ˆ one can always construct a Hermitean operator A as   1 † ˆ ˆ ˆ A = B + B (340) 2 since Aˆ†

  1 † † † ˆ ˆ = B + (B ) 2   1 † ˆ ˆ = B + B 2 = Aˆ

(341)

ˆ The operator Aˆ is the Hermitean part of the arbitrary operator B. ——————————————————————————————————

90

3.4.32

Exercise 26

ˆr = − i ∂ Construct the Hermitean operators corresponding to the operators B ∂r ∂ ˆθ = − i , expressed in spherical polar coordinates r, θ and ϕ. and B ∂θ ——————————————————————————————————

3.4.33

Solution 26

ˆr is found by considering the matrix elements and The Hermitean conjugate of B by integrating by parts with respect to r, Z ∞ Z π Z 2π ∂ − i dr r2 dθ sin θ dϕ Φ∗ (r, θ, ϕ) Ψ(r, θ, ϕ) ∂r 0 0 0  Z ∞ ∗ Z π Z 2π ∂ 2 ∗ = + i dr dθ sin θ dϕ Ψ (r, θ, ϕ) r Φ(r, θ, ϕ) ∂r 0 0 0 (342) where we have assumed that the wave function satisfies the boundary conditions lim

r → ∞

Ψ(r, θ, ϕ) → 0

(343)

ˆr† , is given by Thus, the Hermitean conjugate, B ˆr† = − i 1 ∂ r2 B r2 ∂r

(344)

ˆ From this we can construct a Hermitean operator as the Hermitean part of B, via   i ∂ 1 ∂ 2 ˆ Ar = − + 2 r 2 ∂r r ∂r   ∂ 1 = −i + ∂r r   1 ∂ = −i r (345) r ∂r ˆθ is found from Similarly, the Hermitean conjugate of B Z ∞ Z π Z 2π ∂ − i dr r2 dθ sin θ dϕ Φ∗ (r, θ, ϕ) Ψ(r, θ, ϕ) ∂θ 0 0 0  Z ∞ ∗ Z π Z 2π ∂ = + i dr r2 dθ dϕ Ψ∗ (r, θ, ϕ) sin θ Φ(r, θ, ϕ) ∂θ 0 0 0 (346)

91

The boundary term vanishes at θ = 0 and θ = π due to the presence of the ˆ † is given by sin θ term. Thus, B θ 1 ∂ sin θ sin θ ∂θ a Hermitean operator Aˆθ via   i ∂ 1 ∂ − + sin θ 2 ∂θ sin θ ∂θ   ∂ cos θ −i + ∂θ 2 sin θ   1 1 ∂ 2 θ −i sin 1 ∂θ sin 2 θ

ˆ† = − i B θ From this we can construct Aˆθ

= = =

(347)

(348)

——————————————————————————————————

3.4.34

Eigenvalues and Eigenfunctions of Hermitean Operators

Hermitean operators have eigenvalues and eigenfunctions that have three important properties. These are : (i) The eigenvalues are real numbers. (ii) Any two eigenfunctions corresponding to different eigenvalues have matrix elements which are zero. Wavefunctions that have this property are said to be orthogonal. (iii) Any wave function can be expanded in terms of a complete set of the eigenfunctions of a Hermitean operator. We shall now prove the first two statements. Consider a set of eigenfunctions ˆ These satisfy the eigenvalue equation of the operator A. Aˆ φa (r) = a φa (r)

(349)

Then take the matrix elements with an eigenfunction corresponding to the eigenvalue b, Z Z ∗ 3 ˆ d3 r φ∗b (r) φa (r) (350) d r φb (r) A φa (r) = a but since Aˆ is Hermitean ∗ d3 r φ∗a (r) Aˆ† φb (r) ∗  Z 3 ∗ ˆ = d r φa (r) A φb (r)  Z

=

92

(351)

However, φb (r) is an eigenstate of Aˆ with eigenvalue b, so  Z ∗ = d3 r φ∗a (r) b φb (r) Z = b∗ d3 r φ∗b (r) φa (r)

(352)

Thus, we find the equality ( a − b∗ )

 Z

d3 r φ∗b (r) φa (r)

 = 0

(353)

Thus, if the two eigenvalues are the same, b = a, and φa (r) is normalized to unity, one must have a = b∗ = a∗ . Hence, we have proved that the eigenvalues of a Hermitean operator are real. If the eigenvalues are different, b 6= a, one then has Z d3 r φ∗b (r) φa (r) = 0

(354)

Thus, eigenfunctions of the Hermitean operator corresponding to different eigenvalues are orthogonal. If an eigenvalue is degenerate, then the eigenfunctions corresponding to this eigenvalue are not necessarily orthogonal. However, by choosing appropriate linear combinations of these eigenfunctions one can construct a set of eigenfunctions which are orthogonal. This procedure is known as Gram-Schmidt orthogonalization. We shall implicitly assume that this procedure has always been performed on any set of eigenfunctions. We shall delay the proof of the expansion of an arbitrary wave function till later. However, we shall show how the orthonormality properties of the eigenfunctions allow a simple evaluation of the expansion coefficients Cn of an arbitrary wave function X Cn φn (r) (355) Ψ(r) = n

Taking the matrix elements with the eigenfunction φm (r), one finds Z Z X Cn d3 r φ∗m (r) φn (r) d3 r φ∗m (r) Ψ(r) = n

=

X

Cn δn,m

n

= Cm 93

(356)

Therefore, the expansion of an arbitrary wave function Ψ(r) can be written in the form X Z Ψ(r) = d3 r0 φ∗n (r0 ) Ψ(r0 ) φn (r) (357) n

After re-arranging, one recognizes that the sum over the eigenfunctions has the same property as the Dirac delta function. Thus, we have found the completeness condition X φ∗n (r0 ) φn (r) = δ 3 ( r − r0 ) (358) n

This condition has to be satisfied if any wave function can be expanded in terms of a complete set of eigenfunctions. On going back to the example of momentum eigenfunctions and wave packets, one recognizes that the sum over the eigenvalues n is to be replaced by an integral over the continuous eigenvalues k. Furthermore, the completeness condition is just the representation of the Dirac delta function given by eqn(132). In the case of degenerate eigenvalues, an appropriate orthonormal set of eigenfunctions should be constructed. ——————————————————————————————————

3.4.35

Exercise 27

Express the x, y and z components of the orbital angular momentum operˆ in terms of the spherical polar coordinate variables (r,θ,ϕ). Find all ator L the eigenfunctions and eigenvalues of the z component of the (Hermitean) orˆ z . Show that they form an orthonormal set. bital angular momentum operator L By using the theorems of discrete Fourier transforms for fixed r and θ, show that any wave function can be expanded in terms of the complete set of eigenˆ z , φm (ϕ) as functions of L X Ψ(r, θ, ϕ) = Φm (r, θ) φm (ϕ) (359) m

where the expansion coefficients Φm (r, θ) are functions of r and θ. ——————————————————————————————————

3.4.36

Solution 27

Since the momentum is given by   ∂ 1 ∂ 1 ∂ p = −i¯ h eˆr + eˆθ + eˆϕ ∂r r ∂θ r sin θ ∂ϕ

94

(360)

ˆ = r ∧ pˆ, one finds that the angular momentum and r = r eˆr then with L solely has eˆθ and eˆϕ components. The total orbital angular momentum can then be resolved into the Cartesian components as   ∂ ∂ Lx = − i ¯ h − sin ϕ − cos ϕ cot θ ∂θ ∂ϕ   ∂ ∂ Ly = − i ¯ h + cos ϕ − sin ϕ cot θ ∂θ ∂ϕ ∂ Lz = − i ¯h (361) ∂ϕ The eigenfunctions of the z component of the orbital angular momentum operator are  1   1 2 φm (ϕ) = exp i m ϕ (362) 2π where, because of the condition that φm (ϕ) is single valued φm ( ϕ ) = φm ( ϕ + 2 π )

(363)

then one finds exp[ i m 2 π ] = 1 which implies that m can only have integer values. Since these eigenfunctions only involve ϕ, and not r and θ, the normalization is only due to the integration over ϕ between 0 and 2 π. Any function of ϕ which has period 2 π can be expanded in terms of this set of orthonormal eigenfunctions. ——————————————————————————————————

3.4.37

Exercise 28

ˆ ]] Find the expectation value of the nested commutation relation [ x ˆ, [x ˆ, H ˆ between the Hermitean operators x ˆ and the energy H between the lowest energy eigenfunctions φ0 , i.e., evaluate Z ˆ ] ] φ0 (r) d3 r φ∗0 (r) [ x , [ x , H (364) and then use the completeness relation to prove the Thomas-Reiche-Kuhn sum rule 2 X Z ¯h2 3 ∗ d r φ0 (r) x φn (r) ( En − E0 ) = + (365) 2m n

R which relates the intensities of optical dipole allowed transitions | d3 r φ∗0 (r) x φn (r) |2 and the energies of the photon involved in the transitions ( En − E0 ). If all these quantities are measured experimentally, multiplied and summed over all possible transitions, one should obtain unity. If the sum is different from unity

95

the experiment has failed to identify some transitions or the experimental apparatus has not been calibrated correctly. ——————————————————————————————————

3.4.38

Solution 28

Starting from the commutation relation ¯2 ˆ ]] = − h [x, [x, H m   2 h ¯ ˆ + H ˆ x2 − 2 x H ˆ x x2 H = − m

(366)

and taking the matrix elements between Ψ∗0 (x) and Ψ0 (x), one has   Z +∞ ˆ + H ˆ x2 − 2 x H ˆ x Ψ0 (x) dx Ψ∗0 (x) x2 H − ∞

¯2 h = − m = −

Z

+ ∞

dx Ψ∗0 (x) Ψ0 (x)

− ∞

¯2 h m

(367)

since the wave functions are normalized to unity. The right hand side of the commutator consists of four terms. The first and second term will turn out to be identical, although it may not seem obvious at first. The third and fourth are obviously identical and have already been combined and these give rise to the term which contains a factor of 2. We shall now examine the first term which is given by Z +∞ ˆ Ψ0 (x) dx Ψ∗0 (x) x x H

(368)

− ∞

and introduce a factor of 1 between the two x’s. This 1 can be written as an integral over a Dirac delta function δ(x − x0 ) with respect to x0 . Thus, we have Z

+ ∞

Z

+ ∞

dx − ∞

ˆ Ψ0 (x0 ) dx0 Ψ∗0 (x) x δ( x − x0 ) x0 H

(369)

− ∞

where everything to the right of the delta function has been expressed in terms of the primed variable x0 . Since Ψ0 (x0 ) is an energy eigenfunction it satisfies ˆ Ψ0 (x0 ) = E0 Ψ0 (x0 ) H

96

(370)

This can be substituted in the expression for the term we are trying to evaluate, to yield Z +∞ Z +∞ = dx dx0 Ψ∗0 (x) x δ( x − x0 ) x0 E0 Ψ0 (x0 ) (371) − ∞

− ∞

On writing the delta function in terms of the completeness relation X δ( x − x0 ) = Ψn (x) Ψ∗n (x0 )

(372)

n

one obtains Z = E0

+ ∞

Z

− ∞

= E0 = E0 = E0

+ ∞

dx0 Ψ∗0 (x) x Ψn (x) Ψ∗n (x0 ) x0 Ψ0 (x0 )

− ∞

+ ∞

dx Ψ∗0 (x) x Ψn (x)

− ∞

Z

+ ∞

dx0 Ψ∗n (x0 ) x0 Ψ0 (x0 )

− ∞

+ ∞

dx Ψ∗0 (x) x Ψn (x)

 Z

− ∞

n

X Z = E0

+ ∞

− ∞

Ψn (x) Ψ∗n (x0 ) x0 Ψ0 (x0 )

+ ∞

dx

X  Z

n

X n

Z

− ∞

X Z n

dx0 Ψ∗0 (x) x

− ∞

X Z n

+ ∞

dx

+ ∞

dx0 Ψ∗n (x0 ) x0 Ψ0 (x0 )



− ∞

2 dx Ψ∗0 (x) x Ψn (x)

(373)

where we have interchanged the order of summation and integrations. The second term can be immediately relate it to the first term. The second term is given by Z +∞ ˆ x x Ψ0 (x) dx Ψ∗0 (x) H (374) − ∞

ˆ x x is just x x H. ˆ Let us note that the Hermitean conjugate of the operator H Using the definition of the Hermitean conjugate of an operator, one finds  Z +∞ ∗ Z +∞ ˆ x x Ψ0 (x) = ˆ Ψ0 (x) dx Ψ∗0 (x) H dx Ψ∗0 (x) x x H (375) − ∞

− ∞

Thus, the second term is just the complex conjugate of the first term, which we have seen is real. The first two terms are thus identical. The remaining terms are evaluated in the same way as the first term. We start with the expression Z +∞ ˆ x Ψ0 (x) dx Ψ∗0 (x) x H (376) − ∞

97

ˆ and an x. This 1 can be written as and introduce a factor of 1 between the H an integral over a Dirac delta function δ(x − x0 ) with respect to x0 . Thus, we have Z +∞ Z +∞ ˆ δ( x − x0 ) x0 Ψ0 (x0 ) dx dx0 Ψ∗0 (x) x H (377) − ∞

− ∞

where everything to the right of the delta function has been expressed in terms of the primed variable x0 . On writing the delta function in terms of the completeness relation X δ( x − x0 ) = Ψn (x) Ψ∗n (x0 ) (378) n

one obtains Z +∞ Z = dx − ∞

+ ∞

ˆ dx0 Ψ∗0 (x) x H

X

− ∞

Ψn (x) Ψ∗n (x0 ) x0 Ψ0 (x0 )

n

(379) 0

Since Ψn (x ) is an energy eigenfunction it satisfies ˆ Ψn (x0 ) = En Ψn (x0 ) H

(380)

This can be substituted in the expression for the term we are trying to evaluate, to yield Z +∞ Z +∞ X ˆ = dx dx0 Ψ∗0 (x) x H Ψn (x) Ψ∗n (x0 ) x0 Ψ0 (x0 ) − ∞

Z

− ∞

+ ∞

=

Z

X

X

= =

n

+ ∞

dx0 Ψ∗0 (x) x Ψn (x) Ψ∗n (x0 ) x0 Ψ0 (x0 )

dx − ∞

+ ∞

En

dx

Ψ∗0 (x)

Z

 Z

dx0 Ψ∗n (x0 ) x0 Ψ0 (x0 )

− ∞

+ ∞

En

+ ∞

x Ψn (x)

− ∞

dx

Ψ∗0 (x)

 Z

Z En

+ ∞

x Ψn (x)

− ∞

n

X

Z

− ∞

Z

En Ψn (x) Ψ∗n (x0 ) x0 Ψ0 (x0 )

n

+ ∞

En

n

X

X

− ∞

Z

n

=

dx0 Ψ∗0 (x) x

dx − ∞

=

n

+ ∞

0

dx

Ψ∗n (x0 )

0

0



x Ψ0 (x )

− ∞

+ ∞

dx

Ψ∗0 (x)

− ∞

2 x Ψn (x)

(381)

where we have interchanged the order of summation and integrations. Combining the four term leads to the equation Z +∞ X 2 ( E0 − En ) dx Ψ∗0 (x) x Ψn (x) n

− ∞

2 2 = − ¯h m

(382)

which was to be proved. ——————————————————————————————————

98

3.4.39

Exercise 29

Evaluate all the 6 commutation relations between the various pairs of the difˆ ferent Cartesian components of the orbital angular momentum operators L. ——————————————————————————————————

3.4.40

Solution 29

The components of the angular momentum are given by ˆx L ˆy L

= y pˆz − z pˆy

ˆz L

= x pˆy − y pˆx

= z pˆx − x pˆz (383)

The commutators can be evaluated by expanding them ˆx , L ˆy ] [L

ˆ y ] − [ z pˆy , L ˆy ] = [ y pˆz , L = [ y pˆz , z pˆx ] + [ z pˆy , x pˆz ] = − i ¯h y pˆx + i ¯h x pˆy ˆz = + i ¯h L

(384)

so one has the result ˆx , L ˆ y ] = i ¯h L ˆz [L

(385)

The set of commutators can be summarized as ˆi , L ˆ j ] = i ¯h i,j,k L ˆk [L

(386)

where i,j,k is the antisymmetric Levi-Civita symbol, defined as i,j,k = 1

(387)

if (i, j, k) is an even permutation of (x, y, z), and i,j,k = − 1

(388)

if (i, j, k) is an odd permutation of (x, y, z) or i,j,k is zero otherwise. ——————————————————————————————————

3.4.41

Exercise 30

In our discussion of the classical mechanics of a charged particle in an electromagnetic field, we have seen that the momentum of the particle is changed due

99

to the presence of the vector potential A(r, t) and that the Hamiltonian is of the form H =

p2 q q2 − p . A(r, t) + A2 (r, t) + q φ(r, t) 2m mc 2 m c2

(389)

In the radiation or Coulomb gauge, defined by the condition ∇ . A(r, t) = 0, electromagnetic waves are solely represented by the vector potential. Thus, to lowest order in A, light couples to the particle via the interaction Hint   ˆ int = − q H pˆ . A(ˆ r, t) + A(ˆ r, t) . pˆ (390) 2mc Show that the momentum operator can be expressed as a commutator of the position operator and the Hamiltonian for a charged particle in the scalar potential ˆ 0, represented by the Hamiltonian H −i

¯ h ˆ 0 , rˆ ] pˆ = [ H m

(391)

Hence, by using the energy eigenfunctions Ψn (r) and eigenvalues En of the ˆ 0 , show that if A(r, t) only varies over long length unperturbed Hamiltonian H scales, i.e. A(r, t) ' A(t), then  Z Z im 3 ∗ d r Ψn (r) pˆ . A(r, t) Ψm (r) = En − Em d3 r Ψ∗n (r) rˆ . A(r, t) Ψm (r) ¯h (392) This provides the usual basis for discussing optical transitions involving the absorption or emission of light in terms of dipole transitions, when the wave length of light is much longer than the scale associated with the atomic or electronic structure. ——————————————————————————————————

3.4.42

Solution 30

The matrix elements of the interaction term can be expressed as Z q − d3 r φ∗m (r) pˆ . A φn (r) mc

(393)

where we are using the radiation gauge ∇.A = 0

(394)

in which the vector potential is transverse to the direction of propagation. We can substitute the expression for pˆ given by pˆ = − i

m ˆ0 ] [r, H ¯h 100

(395)

in the matrix elements, to find Z q ˆ 0 ] φn (r) i d3 r φ∗m (r) A . [ r , H hc ¯

(396)

The matrix elements of the commutator can be expanded as two terms Z q ˆ 0 φn (r) + i d3 r φ∗m (r) A . r H ¯h c Z q ˆ 0 A . r φn (r) − i d3 r φ∗m (r) H ¯h c

(397)

Now on using the energy eigenvalue equation ˆ 0 φn (r) = En φn (r) H

(398)

and the approximate r independence of the vector potential A, the first term can be evaluated, yielding  Z  q + i En A . d3 r φ∗m (r) r φn (r) hc ¯  Z  q ˆ 0 r φn (r) A. − i d3 r φ∗m (r) H (399) hc ¯ ˆ 0 are related to the complex conjugate of its HerThe matrix elements of r H ˆ mitean conjugate H0 r, thus  Z  ˆ 0 r φn (r) d3 r φ∗m (r) H  Z ∗ ˆ 0 φm (r) = d3 r φ∗n (r) r H  Z ∗ = Em d3 r φ∗n (r) r φm (r) (400) Hence, we find q +i hc ¯



 Z

 E n − Em

A.

3

d r

φ∗m (r)

 r φn (r)

(401)

which can be re-written as q +i ωn,m A . c

 Z

3

d r

φ∗m (r)

 r φn (r)

(402)

where En − En = h ¯ ωn,m is related to the frequency of the photon involved in the transition. The matrix elements of q r can be considered to provide an electric dipole for the transition, and hence the approximation that A is almost r independent is called the dipole approximation. ——————————————————————————————————

101

3.4.43

Exercise 31

Extend the proof of the Thomas-Reiche-Kuhn sum rule to show Z 2   X ¯ 2 k2 h 3 ∗ ( E n − Em ) d r φn (r) exp i k . r φm (r) = 2m n

(403)

——————————————————————————————————

3.4.44

Exercise 32

The scattering of light is encapsulated by the Kramers - Heisenberg formula. This involves the square of matrix elements of the interaction between a charged particle and the field, which is of the order of the fourth power of the vector potential A. The contributions to the matrix elements for this process are second order in the vector potential, and occurs as a combination of processes involving the paramagnetic interaction squared and the lowest order diamagnetic interaction. At low frequencies, one has Rayleigh scattering of light, which involves a partial cancellation between the contribution of the diamagnetic interaction and the process involving the paramagnetic interaction. In this exercise, we shall express the matrix elements of the diamagnetic interaction in a form similar to that of the second order process involving the paramagnetic interaction, as needed to demonstrate the partial cancellation. The diamagnetic interaction term can be written as Hdia = +

q2 A(r, t) . A(r, t) 2 m c2

(404)

The matrix elements of this interaction can be related to the scalar product of two unit vectors eˆα . eˆβ (405) Show that, eˆα . eˆβ

1 X 1 × m m Em − En  Z Z 3 ∗ × d r φm (r) ( pˆ . eˆα ) φn (r) d3 r0 φ∗n (r0 ) ( pˆ . eˆβ ) φm (r0 )  Z Z ∗ 3 0 ∗ 0 3 0 d r φn (r ) ( pˆ . eˆα ) φm (r ) + d r φm (r) ( pˆ . eˆβ ) φn (r) =

(406) where φm (r) are energy eigenstates with eigenvalues Em . ——————————————————————————————————

102

3.4.45

Solution 32

We shall first write eˆα . eˆβ

=

X

( eˆα . eˆi ) δi,j ( eˆj . eˆβ )

i,j

=

X

( eˆα . eˆi )

i,j

= eˆα .

[ pˆi , xj ] ( eˆj . eˆβ ) i ¯h

[p ˆ, r] . eˆβ i ¯h

(407)

It shall be understood that the scalar product of the operator p ˆ is taken with eˆα , in both terms of the commutator. On taking the matrix elements between φ∗n (r) and φn (r) and inserting the completeness relation, we have Z [p ˆ, r] eˆα . eˆβ = eˆα . d3 r φ∗n (r) φn (r) . eˆβ − i ¯h Z i = eˆα . d3 r φ∗n (r) p ˆ r φn (r) . eˆβ h ¯ Z i − eˆβ . d3 r φ∗n (r) r pˆ φn (r) . eˆα h ¯ Z Z i 3 ∗ = eˆα . d r φn (r) p ˆ d3 r0 δ( r − r0 ) r0 φn (r0 ) . eˆβ h ¯ Z Z i − eˆβ . d3 r φ∗n (r) r d3 r0 δ( r − r0 ) pˆ0 φn (r0 ) . eˆα h ¯  Z Z i X = eˆα . d3 r φ∗n (r) p ˆ φm (r) d3 r0 φ∗m (r0 ) r0 φn (r0 ) . eˆβ h m ¯  Z Z 0 3 ∗ 3 0 ∗ 0 0 − eˆβ . d r φn (r) r φm (r) d r φm (r ) pˆ φn (r ) . eˆα (408) On using the equality ˆ , r ] φn (r) ( Em − En ) φ∗m (r) r φn (r) = φ∗m (r) [ H ¯h ∗ = −i φ (r) pˆ φn (r) m m one obtains the result   Z 1 X 1 eˆα . eˆβ = ˆ ) φ∗m (r) d3 r φ∗n (r) ( eˆα . p m m Em − En  Z 0 3 0 0 0 × d r φm (r ) ( pˆ . eˆβ ) φn (r )    Z 1 X 1 ˆ ) φ∗m (r) + d3 r φ∗n (r) ( eˆβ . p m m Em − E n 103

(409)

Z

d3 r0 φm (r0 ) ( pˆ0 . eˆα ) φn (r0 )

×

 (410)

as was to be proved. ——————————————————————————————————

3.4.46

Hermitean Operators and Physical Measurements

A physical measurements of a physical quantity A can result in a set of real numbers a that are the possible results of the measurement. It is reasonable to ˆ where represent the operator corresponding to A as a Hermitean operator A, the eigenvalues correspond to the possible set of results a. Given a state represented by Ψ(r) which has the expansion in terms of the eigenfunctions of Aˆ X Ψ(r) = Ca φa (r) (411) a

then the probability of finding a specific result a in the measurement of A is given by P (a) = | Ca |2 (412) The orthonormality of the set of eigenfunctions ensures that the probability distribution P (a) is properly normalized. Z d3 r | Ψ(r) |2 = 1 X 2 Z = d3 r Ca φa (r) a X ∗ = Cb Ca δa,b a,b

=

X

| Ca |2

(413)

a

Thus, the sum over probabilities P (a) is also normalized to unity. The moments of the probability distribution are given by the matrix elements of powers of the operator Z n A = d3 r Ψ∗ (r) Aˆn Ψ(r) Z X X ∗ ∗ ˆn = d3 r Cm Φm (r) Cm 0 Φm0 (r) A m0

Z =

d3 r

X

m ∗ ∗ Cm 0 Φm0 (r)

m0

X m

104

Cm anm Φm (r)

=

X

=

X

∗ n Cm 0 Cm δm,m0 am

m,m0

| Cm |2 anm

(414)

m

and in particular the average value of A is given by the matrix elements Z A = d3 r Ψ∗ (r) Aˆ Ψ(r) (415) This is often called the expectation value of Aˆ in the state Ψ(r). ——————————————————————————————————

3.4.47

Exercise 33

Find the average values of the energy and the momentum of a particle in the state represented by the wave function   2 Ψ(r) = C exp − α ( r − r0 ) (416) where α2 =

m ω 2 h ¯ ,

and the Hamiltonian is given by ˆ = H

pˆ2 m ω 2 r2 + 2m 2

(417)

Also find the mean squared deviation of the (vector) position operator. ——————————————————————————————————

3.4.48

Solution 33

The magnitude of the normalization constant is found from the condition   Z 1 = | C |2 d3 r exp − 2 α ( r − r0 )2 (418) The integral is evaluated to yield 1 = | C |2



π 2α

 32 (419)

Thus, we have the normalization  |C | =

105

2α π

 34 (420)

up to an arbitrary choice of phase. The average value of the energy is given by the expectation value of the Hamiltonian Z ˆ Ψ(r) E = d3 r Ψ∗ (r) H  pˆ2 m ω 2 r2 + Ψ(r) 2m 2   3 Z  2   ¯h m ω 2 r2 2α 2 2 3 2 × + = d r 6 α − 4 α ( r − r0 ) 2 π 2m   × exp − 2 α ( r − r0 )2 (421) Z

=

d3 r Ψ∗ (r)



The integral is evaluated as E

= =

3 ¯h2 α 3 m ω2 3 ¯h2 α m ω 2 r20 − + + 2 2m 8α m 2 2 2 2 3 ¯h α 3mω m ω r0 + + 2 2m 8α

(422)

where the first term is the energy of a displaced classical harmonic oscillator. On substituting α, one finds E =

3 m ω 2 r20 + ¯h ω 2 2

(423)

where the last term represents the zero-point energy for the three independent degrees of freedom. The average value of the momentum is zero, due to the fact that the wave function has no spatially varying phase. Hence, the integrand of the expectation value of p ˆ is antisymmetric around r0   Z p = 2i¯ h α | C |2 d3 r ( r − r0 ) exp − 2 α ( r − r0 )2 (424) Thus, the expectation value of the momentum vanishes. The average value of r2 is given by r2 = r20 +

3 4α

(425)

Hence, the mean squared deviation is given by ∆r2rms

= = 106

3 4α 3 ¯h 2mω

(426)

which corresponds to the sum of the mean square displacements along the three Cartesian axes. ——————————————————————————————————

3.4.49

Exercise 34

Find the average value of the z component of the orbital angular momentum and square of the z component of the angular momentum for the following three wave functions Ψ1a (r) Ψ1b (r) Ψ1c (r)

= C1 r f (r) sin θ cos ϕ = C1 r f (r) sin θ sin ϕ = C1 r f (r) cos θ

(427)

Do you need to evaluate the radial part of the integrations? ——————————————————————————————————

3.4.50

Solution 34

The operator corresponding to the z component of the angular momentum is given by ˆ z = − i ¯h ∂ L (428) ∂ϕ Thus, the expectation value of the angular momentum in a state Ψ(r, θ, ϕ) is given by the expression Z ∞ Z π Z 2π 2 ˆ z Ψ(r, θ, ϕ) Lz = dr r dθ dϕ Ψ∗ (r, θ, ϕ) L 0

0

Z = −i¯ h



dr r 0

2

Z

0 π

Z dθ

0

0

2 π

dϕ Ψ∗ (r, θ, ϕ)

∂ Ψ(r, θ, ϕ) ∂ϕ (429)

The expectation value for the first two states involve the integration Z 2π dϕ sin ϕ cos ϕ = 0

(430)

0

whereas the third state is an eigenstate of the z component of the angular moˆ z is zero mentum operator with eigenvalue zero. Hence, the average value of L in all three states.

107

The mean squared value of the z component of the angular momentum is given by Z ∞ Z π Z 2π ˆ 2z Ψ(r, θ, ϕ) L2z = dr r2 dθ dϕ Ψ∗ (r, θ, ϕ) L 0

0

= −¯ h

2

Z

0



dr r

2

π

Z

0

Z

2 π

dθ 0

0

In the first two states one has the factors Z 2π ∂2 2 −¯ h cos ϕ dϕ cos ϕ ∂ϕ2 0 −h ¯2

Z

2 π

dϕ sin ϕ 0

dϕ Ψ∗ (r, θ, ϕ)

∂2 sin ϕ ∂ϕ2

= h ¯

2

Z

∂2 Ψ(r, θ, ϕ) ∂ϕ2 (431)

2 π

dϕ cos2 ϕ

0

= h ¯2 π Z = h ¯2

2 π

dϕ sin2 ϕ

0

= h ¯2 π

(432)

and the corresponding factor for the third state is identically zero as it is an eigenstate with zero eigenvalue. One recognizes that the expectation values are proportional to the normalization integral and, as the state are properly normalized, one has L2z = h ¯2 (433) for the first two states and the expectation value is zero for the last state. ——————————————————————————————————

3.4.51

Exercise 35

Find the average values of the x, y and z component of the orbital angular momentum and square of the component of the angular momentum for the following three wave functions Ψ1a (r) Ψ1b (r) Ψ1c (r)

= C1 ( x + i y ) f (r) = C1 ( x − i y ) f (r) = C1 z f (r)

(434)

where f (r) is a function of the radial distance, defined via r2 = x2 + y 2 + z 2 . What can one conclude about the magnitude of the angular momentum of these examples? ——————————————————————————————————

108

3.4.52

Solution 35

We shall first note that the components of the angular momentum operator are   ˆ x = − i ¯h y ∂ − z ∂ L ∂z ∂y   ˆ y = − i ¯h z ∂ − x ∂ L ∂x ∂z   ˆ z = − i ¯h x ∂ − y ∂ L (435) ∂y ∂x and that any radially symmetric function f (r) is an eigenfunction of the components of the angular momentum operator with a zero eigenvalue. For example, ˆ z on an arbitrary radially symmetric on considering the action of the operator L function f (r), one has   ∂f ∂f ˆ Lz f (r) = − i ¯h x − y ∂y ∂x   x ∂f y − y = − i ¯h x r r ∂r = 0 (436) as long as r 6= 0. Likewise, the other components also produce zero when acting on an arbitrary radial function f (r). ˆ z with The three functions given by Ψ1 (r) are seen to be eigenfunctions of L eigenvalues ± ¯ h and 0, since   ˆ z Ψ1a/b (r) = − i ¯h C1 f (r) x ∂ − y ∂ L (x ± iy) ∂y ∂x   = − i ¯h C1 f (r) ± xi − y1 = ±¯ h C1 f (r) ( x ± i y ) = ±h ¯ Ψ1a/b

(437)

while ˆ z Ψ1c (r) L

 = − i ¯h C1 f (r) =

0

∂ ∂ x − y ∂y ∂x

 z (438)

ˆ z and the average value of Lz is equal to Thus, these states are eigenstates of L either ± ¯ h or 0 respectively. Also the average values of L2z are just 1 and 0. ˆ x and L ˆ y in states Ψ1a/b (r) can be found by noting The average value of L that ˆ x Ψ1a/b (r) = ∓ ¯h Ψ1c (r) L ˆ y Ψ1a/b (r) = − i ¯h Ψ1c (r) (439) L 109

and the expectation values of Lx and Ly are zero in these states since the origiˆ z corresponding to nal states and the transformed states are all eigenstates of L different eigenvalues, and the eigenstates of a Hermitean operator with different eigenvalues are orthogonal. ˆ x and L ˆ y in the state Ψ1c (r) can be found by The expectation value of L noting that ˆ x Ψ1c (r) L

= − i ¯h C1 y f (r)   ¯h = − Ψ1a (r) − Ψ1b (r) 2

ˆ y Ψ1c (r) L

= +ih ¯ C1 x f (r)   ¯h = i Ψ1a (r) + Ψ1b (r) 2

(440)

ˆ z . The expectation which are just linear combinations of the eigenstates of L ˆ value is zero as the overlap between eigenstates of Lz with different eigenvalues is zero. We recognize that the set of functions Ψ1a (r), Ψ1b (r) and Ψ1c (r) form a closed set under the angular momentum operators, as they operators acting on any member of the set just transform them into linear combinations of the set. By repeated use of the above relations one can show that Z Z ˆ 2x Ψ1a (r) = − ¯h ˆ x Ψ1c (r) d3 r Ψ∗1a (r) L d3 r Ψ∗1a (r) L   Z ¯h2 = d3 r Ψ∗1a (r) Ψ1a (r) − Ψ1c (r) 2 = and also that Z ˆ 2x Ψ1b (r) d3 r Ψ∗1b (r) L

¯h2 2

(441)

Z = +h ¯ = − =

¯2 h 2

ˆ x Ψ1c (r) d3 r Ψ∗1b (r) L   Z ∗ 3 d r Ψ1b (r) Ψ1a (r) − Ψ1c (r)

¯2 h 2

(442)

ˆ 2 one finds that Likewise, for the expectation values of L y Z Z ˆ 2y Ψ1a (r) = − i ¯h ˆ y Ψ1c (r) d3 r Ψ∗1a (r) L d3 r Ψ∗1a (r) L

110

and also that Z ˆ 2y Ψ1b (r) d3 r Ψ∗1b (r) L

=

¯2 h 2

=

¯2 h 2

Z

3

d r

Ψ∗1a (r)



 Ψ1a (r) + Ψ1c (r) (443)

Z = − i ¯h = − =

¯2 h 2

ˆ y Ψ1c (r) d3 r Ψ∗1b (r) L   Z d3 r Ψ∗1b (r) Ψ1a (r) + Ψ1c (r)

¯2 h 2

(444)

ˆ 2x and L ˆ 2y in the state Ψ1c (r) are found from The expectation value of L Z

3

d r

Ψ∗1c (r)

ˆ 2 Ψ1c (r) L x

  Z ¯ h 3 ∗ ˆ = − d r Ψ1c (r) Lx Ψ1a (r) − Ψ1b (r) 2 Z = h ¯2 d3 r Ψ∗1c (r) Ψ1c (r) = h ¯2

and also that Z ˆ 2y Ψ1c (r) d3 r Ψ∗1c (r) L

(445)

  Z ¯ h 3 ∗ ˆ = +i d r Ψ1c (r) Ly Ψ1a (r) + Ψ1b (r) 2 Z = h ¯2 d3 r Ψ∗1c (r) Ψ1c (r) = h ¯2

(446)

b2 = L ˆ2 + L ˆ2 + L ˆ 2 is 2 ¯h2 for all these states. In fact The average value of L x y z one can show that the states Ψ1a (r), Ψ1b (r), Ψ1c (r) are not only eigenstates of ˆ z but are are all eigenstates of the operator L c2 = L ˆ2 + L ˆ2 + L ˆ 2z L x y

(447)

with eigenvalues of 2 ¯h2 . This operator corresponds to the squared length of the vector angular momentum. These states can be thought of as having angular momentum of the same magnitude, but with different orientations. ——————————————————————————————————

111

3.5

Quantization

We have emphasized that because a physical measurement results in a change of the state of the system, measurements should be represented by operators. We have also asserted that a measurement of a physical quantity A yields a real value a as the result of the measurement, and that immediately after the measurement has been made it is certain that the resulting state corresponds to the state in which A has the definite value a. This means that, immediately after the measurement, the system is in an eigenstate of the operator Aˆ which has an eigenvalue a. Furthermore, we have made a choice that the physical operators should be Hermitean so that we know that the eigenvalues are real. Thus far, we have been mainly examining the mathematics of operators and wave functions using several physical examples. It remains to discuss how these operators are chosen, and what the relationships between them are.

3.5.1

Relations between Physical Operators

A physical operator that appears in Quantum Mechanics is associated with the name of an analogous quantity or variable in Classical Mechanics. That is, we associate a classical variable with a quantum mechanical operator. In the Hamiltonian formulation of classical mechanics, the classical quantities may be expressed in terms of coordinates and momenta and perhaps time, but not in terms of derivatives. These relationships are called constitutive or definitive relationships. In quantum mechanics, the various operators representing physical quantities satisfy the same constitutive relationships as the classical variables do. That is, the classical quantities in a definitive relationship are replaced by the quantum operators. However, the order in which the quantum operators appear in the definition may have to be chosen judiciously, as they may not be commuting operators.

3.5.2

The Correspondence Principle

The relationships between the various quantum operators are chosen to correspond to the relations between variables in classical mechanics, for a specific purpose. Classical Mechanics has a well defined regime in which it provides an excellent and accurate description of nature. This is often called the regime of validity of Classical Mechanics. Quantum Mechanics, if it is to provide a better description of nature, should produce results which are identical to the results of classical mechanics (to within the experimental accuracy) within the regime of validity of classical mechanics. By assuming that the same relationships hold between classical variables and quantum operators, it is ensured that if the quantum operators can, to a good approximation, be replaced by values, i.e. real variables, then Quantum Mechanics will reduce to classical mechanics. We shall examine the conditions that are required for a quantum state to lie within the regime of validity of Classical Mechanics. One condition that is needed to be 112

satisfied, if quantum operators may be approximated by real variables, is that ˆ ] between two operators Aˆ and the value associated with the commutator [ Aˆ , B ˆ should be negligibly small when compared to the products of the values assoB ˆ The values or variables associated with the operators are ciated with Aˆ and B. naturally identified with the expectation values for the quantum state in question. Since, in Hamiltonian mechanics, all quantities are expressed in terms of the canonically conjugate momenta and coordinates qi and pj , the commutators of all operators can be reduced to the non-zero combinations of the fundamental commutator [ qˆi , pˆj ]. The Complementarity Principle associates the value of the fundamental commutators with the corresponding classical Poisson Brackets, up to a constant proportional to h ¯ . If the commutators, hence h ¯ , can be considered to be negligibly small when compared with the average values of pi and qj in the physical state, then Quantum Mechanics will reduce to Classical Mechanics in this limit. The Correspondence Principle can be embodied in the concise statement that Quantum Mechanics should reduce to Classical Mechanics in the limit h ¯ → 0.

3.5.3

The Complementarity Principle

The complementarity principle has the effect that the Poisson Brackets between two classical variables are to be replaced by the commutator between two quantum operators ˆ ] [ Aˆ , B (448) [ A , B ]P B → i ¯h where ¯h is a universal constant. The equivalence can be seen by examining the Poisson Bracket algebra between two products of quantum operators, while respecting the order ˆ1 B ˆ2 ]P B [ Aˆ1 Aˆ2 , B

ˆ1 ]P B B ˆ2 Aˆ2 + [ Aˆ1 , B ˆ1 ]P B B ˆ2 + Aˆ1 [ Aˆ2 , B ˆ1 ]P B Aˆ2 B ˆ2 + = [ Aˆ1 , B ˆ ˆ ˆ + B1 [ A1 , B2 ]P B Aˆ2

=

ˆ1 [ Aˆ1 , B ˆ1 + Aˆ1 B Aˆ1 [ Aˆ2 , ˆ1 + Aˆ1 B

ˆ2 ]P B Aˆ2 B ˆ2 ]P B [ Aˆ2 , B ˆ1 ]P B B ˆ2 B ˆ ˆ [ A2 , B2 ]P B (449)

Thus, on equating these two results we find ˆ1 ]P B [ Aˆ2 , B ˆ2 ] = [ Aˆ2 , B ˆ3 ]P B [ Aˆ1 , B ˆ1 ] [ Aˆ1 , B

(450)

ˆ1 are independent of the pair Aˆ2 and which, since the pair of operators Aˆ1 and B ˆ B2 , leads to the discovery that the quantum Poisson bracket is proportional to the commutator, with a universal constant of proportionality ˆ1 ] = i ¯h [ Aˆ1 , B ˆ1 ]P B [ Aˆ1 , B

(451)

On retaining the classical value of the Poisson bracket between canonically conjugate coordinates qi and momentum pj , one finds that the quantum operators 113

must satisfy the commutation relations [ pˆj , qˆi ] = − i ¯h δj,i [ pˆj , pˆi ] = [ qˆj , qˆi ] = 0

(452)

Having found the commutation relationships between the momenta and coordinate operators, the basis for quantization is complete as any operator can be expressed in terms of the canonical conjugate momenta and coordinates. However, it is usual to find explicit representations for the operators. The procedure of quantization is simplest and free from ambiguities, if one quantizes in a Cartesian coordinate system and then transforms to other coordinate systems. The classical limit is obtained when the operators can be replaced by their eigenvalues and this implies that the commutators are negligible. As the commutation relation between two operators ultimately involves a commutation relation between position and momentum, the commutation relation is proportional to h ¯ . The classical limit is approached when h ¯ is negligible compared with the products of appropriate expectation values in a state.

3.5.4

Coordinate Representation

The coordinate representation is that usually used in Schr¨odinger’s wave mechanics. In this case, the position operator ˆr is diagonal in that it is represented by the vector variable r. The momentum operator p ˆ is represented by a Hermitean first order differential operator. The wave function is a function of the position Ψ(r), as r is diagonal. The momentum operator is represented as the sum of a gradient term and a real function ˆr = r p ˆ = − i ¯h ∇ + ∇ Λ(r)

(453)

where Λ(r) is an arbitrary real function. The gradient of the real function, ∇ Λ, is to be regarded as a function and not as an operator. Usually the arbitrary function is removed from the momentum operator by an appropriate choice of the phase of the wave function. → p ˆ0 = − i ¯h ∇   Λ(r) Ψ(r) → Ψ0 (r) = exp + i Ψ(r) ¯h p ˆ

(454)

This representation is the one we shall mostly use in this class. For example, in generalized coordinate systems, the infinitesimal displacement ds is given in terms of the metric gi,j and the infinitesimal changes in the 114

generalized coordinates dq j via X

ds2 =

gi,j dq i dq j

(455)

i,j

The metric is a symmetric tensor gi,j = gj,i . Since the kinetic energy is defined 2 in terms of the square of the infinitesimal displacement per unit time, ( ds dt ) , the classical Lagrangian is given by m X i L(q, q) ˙ = q˙ gi,j q˙j − V (q) (456) 2 i,j The canonical momenta are defined by the derivatives of the Lagrangian w.r.t. the generalized velocities pi =

∂L = m gi,j q˙j ∂ q˙i

(457)

The associated quantum mechanical operators have to be Hermitean with the inner product Z Y √ dq j g Φ∗ (r) Ψ(r) (458) j

where g is given by the determinant of the covariant metric tensor   g = det gi,j (q)

(459)

The quantum mechanical operators must satisfy the canonical commutation rules [ pˆi , qˆj ] = − i ¯h δij [ pˆi , pˆj ] = 0 [ qˆi , qˆj ] = 0

(460)

In the coordinate representation, the generalized momenta are given by 1

pˆj = − i ¯h g − 4 or

 pˆj = − i ¯h

where Γj = g

− 12

1 ∂ g4 ∂q j

1 ∂ + Γj j ∂q 2 

1 ∂ g2 ∂q j

(461)  (462)

 (463)

The position operators are given by qˆj = q j 115

(464)

As shown by Podolsky8 , the Hamiltonian can be written in the symmetric Hermitean form X 1 1 1 1 ˆ = 1 g − 4 pˆi g 4 g i,j g 4 pˆj g − 4 + V (q) (465) H 2 m i,j −1 where the contra-variant metric tensor is given by g i,j = gi,j .

In planar polar coordinates, (r, ϕ), the infinitesimal length element is given by ds2 = dr2 + r2 dϕ2 Thus, the co-variant metric has the diagonal form   1 0 gi,j = 0 r2

(466)

(467)

and the contra-variant metric is g

i,j

 =

1 0

0

 (468)

1 r2

The infinitesimal volume element is given by √ dr dϕ g = dr dϕ r

(469)

The generalized momenta operators are given by 1

pˆr

= − i ¯h r− 2

pˆϕ

= − i ¯h r− 2

1

∂ 1 r2 ∂r ∂ 1 r2 ∂ϕ

(470)

The Hamiltonian is given by   1 1 1 1 1 1 ˆ = + V (r) r− 2 pˆr r pˆr r− 2 + r− 2 pˆϕ pˆϕ r− 2 H r 2m     h2 ¯ 1 ∂ ∂ 1 ∂2 = − (471) r + 2 + V (r) 2 m r ∂r ∂r r ∂ϕ2

In spherical polar coordinates, (r, θ, ϕ), the infinitesimal length ds is given by ds2 = dr2 + r2 dθ2 + r2 sin2 θ dϕ2 Thus, the co-variant metric has the diagonal  1 0 gi,j =  0 r2 0 0 r2 8 B.

Podolsky, Phys. Rev. 32, 812 (1928).

116

(472)

form  0  0 sin2 θ

(473)

and the contra-variant tensor is given by the inverse matrix   1 0 0  0 g i,j =  0 r12 1 0 0 r2 sin 2θ Thus, the infinitesimal volume element is given by √ dr dθ dϕ g = dr dθ dϕ r2 sin θ

(474)

(475)

and the canonical momentum pˆj are given by pˆr

pˆθ

pˆϕ

√ 1 ∂ r sin θ = − i ¯h √ r sin θ ∂r 1 ∂ = − i ¯h r r ∂r √ 1 ∂ r sin θ = − i ¯h √ r sin θ ∂θ 1 ∂ √ = − i ¯h √ sin θ sin θ ∂θ √ 1 ∂ = − i ¯h √ r sin θ r sin θ ∂ϕ ∂ = − i ¯h ∂ϕ

(476)

Hence, the Hamiltonian has the form   1 1 1 1 1 1 ˆ = H pˆr r2 pˆr + √ pˆθ sin θ pˆθ √ + pˆϕ 2 p ˆ + V (r) ϕ 2m r r r sin2 θ r sin θ r sin θ       h2 ¯ 1 ∂ 1 ∂ ∂ 1 ∂2 2 ∂ = − r + 2 sin θ + 2 + V (r) 2 m r2 ∂r ∂r r sin θ ∂θ ∂θ r sin2 θ ∂ϕ2 (477)

3.5.5

Momentum Representation

The momentum representation is the representation in which the momentum operator is diagonal, and the wave function is given as a function of p, Φ(p). The momentum and position operators are given by p ˆ = p ˆr = + i ¯h ∇p − ∇p Λ(p)

(478)

where the gradients are derivatives w.r.t. p . This representation of the operators satisfies the required commutation relations between the components of rˆ 117

and pˆ . In complete analogy to the real-space representation, the momentum distribution function is given by P (p) d3 p = | Φ(p) |2 d3 p

(479)

and is normalized to unity Z

P (p) d3 p = 1

(480)

The transform between the real space and momentum representation is given by the integral Fourier Transform,  Ψ(r) =

1 2 π ¯h

 32 Z

1 2 π ¯h

 32 Z

d3 p exp

p.r ¯h



+ i

p.r ¯h



− i



Φ(p)

(481)

Ψ(r)

(482)

and its inverse  Φ(p) =

d3 r exp



The components of momentum pˆ and ˆr position operators in the coordinate representation are defined by their action on an arbitrary state Ψ(r), which produces transformed states ψpx and ψx ψpx (r) ψx (r)

= − i ¯h ∇x Ψ(r) = rx Ψ(r)

(483)

The corresponding operators in the momentum representation are found by Fourier transforming these equations. On Fourier Transforming and integrating the first equation by parts, and with appropriate boundary conditions, one finds the expression for the momentum operator in the momentum representation. Likewise, on differentiating the integral Fourier transform with respect to px , one obtains a representation of the position operator ˆrx in momentum space. The resulting equations are φpx (p) = px Φ(p) φx (p) = + i ¯h ∇px Φ(p)

(484)

This confirms the identification of the operators in the momentum representation. —————————————————————————————————— Example

118

The momentum space representation provides a simple way for finding the bound states of a particle moving in one dimension, in the presence of a particular potential V (x) → ∞

for x < 0 2

V (x)



=

e 2x

for x > 0

(485)

Image potential for electrons near a metal surface

V(x) x0 / e

2

1 0.5 0 -0.5 2

-1

V(x) = - e / 2 x

-1.5 -2 -2

0

2

4

6

8

10

x / x0

Figure 21: The one-dimensional image potential V (x) which confines a particle close to an in-penetrable region for x < 0. The bound state energy is marked by a horizontal line. —————————————————————————————————— Solution On multiplying by x, the energy eigenvalue equation becomes   pˆ2 e2 x ˆ − φn (p) = x ˆ En φn (p) 2m 2

(486)

A simplification has occurred for our particular choice of potential V (x), since the energy eigenvalue equation has been transformed so that it is now linear in x ˆ. Therefore,  2  ∂ p e2 i¯ h − En φn (p) = φn (p) (487) ∂p 2 m 2 119

This is a first order differential equation    2   p e2 p ∂φn i ¯h − φn (p) = − i ¯h − En m 2 2m ∂p The equation can be put in the form of an integral   ∂φn  p0 Z p Z p i e2  ∂p0 0 0 m + h ¯ 2 dp = − dp p0 2 φn (p0 ) 0 0 2 m − En

(488)

(489)

which can be evaluated as r   φn (p) p2 i e2 2m p = − ln 1 − − tan−1 √ ln (490) φn (0) 2 m En 2 ¯h − En −2 m En Hence, one finds r     i e2 2m p φn (0) −1 √ exp − tan (491) φn (p) = 2 2 ¯h − En − 2 m En 1 − 2 mp En Thus, φn (p) vanishes as p → ± ∞ and is normalizable. This indicates that ψn (x) is also normalizable for En < 0 and hence satisfies the boundary condition at x → ∞. The energy eigenvalues are found by insisting that the real space wave function ψn (x) satisfy the boundary condition ψn (0) = 0

(492)

at x = 0, since V (0) → ∞. The wave function in real space is given in terms of the momentum space wave function through   Z ∞ 1 px ψn (x) = √ dp φn (p) exp i (493) ¯h 2 π ¯h −∞ so the boundary condition at x = 0 becomes Z ∞ 1 √ dp φn (p) 0 = ψn (0) = 2 π ¯h −∞

(494)

The boundary condition has the explicit form r     Z ∞ 1 φn (0) i e2 2m p −1 0 = √ dp exp − tan √ 2 2 ¯h − En − 2 m En 2π¯ h −∞ 1 − 2 mp En (495)

120

The imaginary part of the integral is odd in p and, thus, vanishes identically. Therefore, the boundary condition is satisfied, if one requires that the real part vanishes    2 r  Z ∞ 1 e φn (0) 2m p −1 cos 0 = √ dp tan √ 2 2 ¯h − En − 2 m En 2π¯ h −∞ 1 − 2 mp En (496) On integrating this, one finds the condition r   π e2 2m sin = 0 (497) 4 ¯h − En Hence, the energy eigenvalues are determined by the boundary condition r 2m π e2 = nπ (498) − En 4 ¯h and are given by En = −

2 m e4 16 ¯h2 n2

(499)

which has the form of a Rydberg series. —————————————————————————————————— Example: A Particle Confined in a Uniform Force Field. Another problem which can be solved in the momentum representation is that of a particle of mass m moving in one dimension in the presence of a potential V (x)

= −F x → ∞

if x < 0 if x > 0

(500)

The potential is depicted in fig(22). A potential of this type confines electrons near the interface of semiconductors with different levels of doping9 . The infinite potential excludes the particle from the region x > 0 and introduces the boundary condition at x = 0 φ(0) = 0 (501) The energy eigenvalue equation becomes −

¯ 2 ∂2φ h − F x φ(x) = E φ(x) 2 m ∂x2

(502)

for x < 0. The classical turning point − a is given by a = 9 T.S.

E F

Rahman, D.L. Mills and P.S. Riseborough, Phys. Rev. B 23, 4081, (1981).

121

(503)

2 2

V(x) [ units of (h F /2m)

1/3

]

Confining Potential 15

En 10

-Fx E1

5

E0 0 -15

-10

-5

0

5

ξ x/ξ

Figure 22: The potential of the uniform force and the infinite potential confines the particle in the region x < 0. The energy eigenvalues En are depicted by the dashed lines. On eliminating the energy in the eigenvalue equation in favor of the turning point, one finds h2 ∂ 2 φ ¯ − F (x + a)φ = 0 (504) − 2 m ∂x2 A characteristic length scale ξ is defined by  ξ =

¯2 h 2mF

 13

Hence, after introducing the dimensionless variable z = the eigenvalue equation for z ξ < a has the form ∂2φ + zφ = 0 ∂z 2

(505) x + a ξ ,

one finds that

(506)

This equation is related to Airy’s equation and has solutions which are linear combinations Ai(−z) and Bi(−z). The function Bi(z) diverges exponentially as z → ∞. Hence, a solution containing Bi(−z) is not a physically acceptable solution in the region x < 0. The function Bi(−z) is discarded, so the solution is of the form   x+a φ(x) = C Ai − (507) ξ

122

for x < 0. The solution has the integral representation     Z ∞ ξ x+a 1 3 3 = dk cos k ξ − k(x + a) Ai − ξ π 0 3

(508)

The asymptotic form of the wave function can be found by the method of stationary phase and is given by C φ(x) ∼ √ π



− 14

x+a ξ

 sin

2 3



x+a ξ

 32

π + 4

 (509)

within the classically accessible region (0 > x > − a), and decays like φ(x) ∼

2

C √

 π

x+a − ξ

− 14

 exp

2 − 3



x+a − ξ

 32  (510)

in the classically forbidden region (x < − a). However, to be an eigenfunction, the wave function must also satisfy the boundary condition of eqn(501) at x = 0. Hence, one requires   an = 0 (511) Ai − ξ The boundary condition determines the allowed values of the turning point a which are labeled by an , or equivalently, by using eqn(503), the boundary condition determines the energy eigenvalue En . The asymptotic large n values of the ratio is given by the formula an = ξ



3π (4n − 1) 8

 23 (512)

This approximate formula is quite accurate even for n = 1, as can be seen from 2 Table(1). For large n, the zeros vary sub-linearly as an ∼ n 3 . Hence, the asymptotic form of the energy eigenvalues is given by  En =

¯2 F 2 h 2m

 13 

3π (4n − 1) 8

 23 (513)

so the level spacing decreases with increasing n. The first few wave functions are shown in fig(23). —————————————————————————————————— 3.5.6

Exercise 36 2

p ˆ Find the momentum representation of the kinetic energy operator Tˆ = 2 m starting from the known expression in the real space or position representation

¯2 h Tˆ = − ∇2 2m 123

(514)

Table 1: The first few zeros of the Airy function. n

an / ξ (exact value)

an / ξ (asymptotic)

1 2 3 4 5 6 7 8 9

2.338 4.088 5.521 6.787 7.944 9.023 10.040 11.009 11.936

2.320 4.082 5.517 6.784 7.942 9.021 10.039 11.008 11.935

The lowest energy eigenfunctions 0.8

φ2(x)

φ3(x)

0.6

φ1(x)

φn(x)

0.4 0.2 0 -0.2 -0.4 -0.6 -10

-8

-6

x/ξ

-4

-2

0

Figure 23: The x dependence of the first few eigenfunctions. The wave functions are non-zero in the region x < 0. The wave functions decay beyond the classical turning points and are zero for x > 0.

——————————————————————————————————

124

3.5.7

Exercise 37

Starting from the real space representation of the pseudo-vector angular moˆ = ˆr ∧ p mentum operator L ˆ, acting on a wave function, find the momentum representation of the operator through integration by parts. ——————————————————————————————————

3.5.8

Exercise 38

Find a momentum space representation for the Hamiltonian of a particle moving in three dimensions, in the presence of an electrostatic potential φ(ˆ r). Also write the explicit expression that occurs when the potential produces a uniform electric field E(r, t) = E 0 . Find the momentum space eigenfunctions for this particular Hamiltonian.

Potential due to a Uniform Force Field V(x)

6

E 0

V(x) = - F x

-6

-12 -4

-2

0

2

4

6

8

10

12

x

Figure 24: The potential V (x) = − F x, which gives rise to a unform force field F . The energy E is denoted by a horizontal line. ——————————————————————————————————

125

Real-space energy eigenfunction of a particle moving in a uniform field Ψ(x)

2

1

0 -4

-2

0

2

4

6

8

10

12

-1

-2

x

Figure 25: The real space wave function Ψ(x) of a particle of mass m and energy E moving in a uniform force field. 3.5.9

Solution 38

The Hamiltonian in the momentum representation has the form ∂ p2 − i ¯h F 2m ∂p

ˆ = H

(515)

The energy eigenvalue equation is  2  p ∂ − i ¯h F Φ(p) = E Φ(p) 2m ∂p

(516)

and has the solution  Φ(p) = C exp



i ¯h F



p3 − pE 6m



whence the real space wave function is found as   Z ∞ 2C p3 p(F x + E) Ψ(x) = √ dp cos − 6 m ¯h F ¯ F h 2π¯ h 0

(517)

(518)

——————————————————————————————————

126

3.5.10

Exercise 39

Find the two lowest energy eigenvalues and eigenfunctions for a one-dimensional harmonic oscillator, in the momentum representation. ——————————————————————————————————

3.5.11

Solution 39

The normalized ground state wave function for the one-dimensional Harmonic oscillator is given by  ψ0 (x) =

mω π ¯h

 14

m ω x2 2 ¯h

 −

exp

 (519)

The momentum space wave function φ0 (p) is given by the Fourier transform   Z ∞   1 px ψ0 (x) (520) φ0 (p) = √ dx exp − i ¯h 2 π ¯h −∞ which is evaluated by completing the square and integrating  φ0 (p) =

1 π ¯h m ω

 14

p2 2 ¯h m ω

 exp



 (521)

The momentum space wave function is properly normalized, thus P (p) = | φ0 (p) |2 is the probability density for finding the particle with momentum p. It is notable that the most probable value of the momentum is zero, as the classical particle would be at rest. The normalized first excited state wave function is given by  ψ1 (x) =

2 √ π

 12 

mω ¯h

 34

 x exp

m ω x2 − 2 ¯h

 (522)

The momentum space wave function is evaluated as  φ1 (p) =

2 √ π

 12 

1 ¯ mω h

 34

 p exp



p2 2 ¯h m ω

 (523)

which is also properly normalized. It should be noted that the most probable values of p is non-zero, as in the excited state the classical particle has finite values of the momentum. ——————————————————————————————————

127

3.5.12

Exercise 40

Prove that the overlap matrix between two different states Ψ(r) and ψ(r) in position representation are related via Z Z 3 ∗ d r Ψ (r) ψ(r) = d3 p Φ∗ (p) φ(p) (524) to the corresponding wave functions, Φ(p) and ψ(p), in the momentum space representation. —————————————————————————————————— All observable physical quantities can be expressed in terms of matrix elements of operators. ——————————————————————————————————

3.5.13

Solution 40

On evaluating the integral Z by using the transform  φ(p) =

1 2 π ¯h

d3 p Φ∗ (p) φ(p)

 32 Z

3

d r exp



p.r − i ¯h

(525)

 ψ(r)

and the complex conjugate   32 Z   p . r0 1 Φ∗ (p) = Ψ∗ (r0 ) d3 r0 exp + i 2π¯ h ¯ h

(526)

(527)

one has   3 Z  Z Z ( r0 − r ) . p 1 3 3 0 3 Ψ∗ (r0 ) ψ(r) (528) d r d r d p exp i ¯h 2 π ¯h Since the integral over all space leads to the three-dimensional Dirac delta function   3 Z  ( r0 − r ) . p 1 (529) = δ 3 ( r0 − r ) d3 p exp i ¯h 2π¯ h the overlap between the states is given by Z Z Z d3 p Φ∗ (p) φ(p) = d3 r d3 r0 Ψ∗ (r0 ) δ( r0 − r ) ψ(r) Z = d3 r Ψ∗ (r) ψ(r) (530) 128

as was to be shown. ——————————————————————————————————

3.5.14

Exercise 41

Consider the following matrix elements in the position representation   Z d3 r Ψ∗ (r) − i ¯h ∇ ψ(r) and

Z

d3 r Ψ∗ (r) r ψ(r)

(531)

(532)

Find the equivalent expressions in the momentum space representation. ——————————————————————————————————

3.5.15

Solution 41

Starting with the expression   Z d3 r Ψ∗ (r) − i ¯h ∇ ψ(r)

(533)

and substituting the expressions for the Fourier Transformed wave functions Z Z Z p0 . r p.r d3 r 3 0 ∗ 0 − i¯ h d p exp[ − i ] Φ (p ) ∇ d3 p exp[ + i ] φ(p) 3 (2π¯ h) ¯h ¯h Z Z Z p0 . r p.r d3 r 3 0 ∗ 0 = d p exp[ − i ] Φ (p ) d3 p p exp[ + i ] φ(p) 3 (2π¯ h) ¯h ¯h (534) Then on interchanging the orders of integration and on recognizing the expression for the Dirac delta function   Z ( p − p0 ) . r d3 r δ 3 ( p − p0 ) = (535) exp i ¯h ( 2 π ¯h )3 one has Z

d3 r Ψ∗ (r)



 − i ¯h

Z ∇ ψ(r) =

129

d3 p Φ∗ (p) p φ(p)

(536)

The procedure for evaluating the matrix elements of the position operator is quite analogous. Starting with the expression Z d3 r Ψ∗ (r) r ∇ ψ(r) (537) and substituting the expressions for the Fourier Transformed wave functions Z Z Z p0 . r p.r d3 r 3 0 ∗ 0 d p exp[ − i ] φ(p) ] Φ (p ) r d3 p exp[ + i (2π¯ h)3 ¯h ¯h Z Z Z p0 . r p.r d3 r 3 0 ∗ 0 = −i¯ h d p exp[ − i ] Φ (p ) d3 p φ(p) ∇p exp[ + i ] (2π¯ h)3 ¯h ¯ h (538) Integrating by parts with respect to p, and as the boundary term is zero, we find Z Z Z p.r p0 . r d3 r ∗ 0 3 0 ] ∇p φ(p) ] Φ (p ) d3 p exp[ + i d p exp[ − i = +i¯ h 3 ¯h ¯ h (2π¯ h) (539) Then on interchanging the orders of integration and on recognizing the expression for the Dirac delta function   Z ( p − p0 ) . r d3 r δ 3 ( p − p0 ) = exp i (540) ( 2 π ¯h )3 ¯h one has

Z

3

Z



d r Ψ (r) r ψ(r) = + i ¯h

d3 p Φ∗ (p) ∇p φ(p)

(541)

which completes the solution. We see that the matrix elements between the wave functions and operators in the real space representation have identical values to the equivalent expressions in the momentum representation. ——————————————————————————————————

3.5.16

Commuting Operators and Compatibility

ˆ commute, [ Aˆ , B ˆ ] = 0, these operators are comIf two operators Aˆ and B patible in the sense that it is possible to find functions that are simultaneously ˆ eigenfunctions of Aˆ and B. The proof is simple if the eigenfunctions of the operator Aˆ are non-degenerate. In this case, the eigenfunctions of Aˆ satisfy Aˆ φa (r) = a φa (r) 130

(542)

ˆ one has Taking the matrix elements of the commutator of Aˆ and B, Z ˆ ] φa (r) = 0 d3 r φ∗a0 (r) [ Aˆ , B

(543)

which is evaluated as ( a0 − a )

Z

ˆ φa (r) = 0 d3 r φ∗a0 (r) B

(544)

ˆ only has non-zero matrix elements between states of the Thus, the operator B same eigenvalue a = a0 . Since by assumption the eigenfunctions of Aˆ are nondegenerate and form a complete set, the eigenfunctions of Aˆ are simultaneously ˆ eigenfunctions of B. The converse is that, if there are simultaneous eigenfunctions of the two ˆ then the operators commute. By definition one has, operators Aˆ and B, Aˆ φn (r) = an φn (r) ˆ φn (r) = bn φn (r) B

(545)

Then, the action of the commutator on the wave function is zero as ˆ φn (r) Aˆ B

= = = =

Aˆ bn φn (r) an bn φn (r) ˆ an φn (r) B ˆ Aˆ φn (r) B

(546)

and since an arbitrary wave function Ψ(r) can be expanded in terms of φn (r) the commutator is identically zero. Clearly an example of two commuting operators is given by any operator Aˆ ˆ with a Taylor expansion, and any function f (A) ˆ ] = 0 [ Aˆ , f (A)

(547)

ˆ do not form an independent set, as one can be The operators Aˆ and f (A) ˆ expressed in terms of the other members of the set. Alternatively if Aˆ and B commute then ˆ , f (B) ˆ ] = 0 [ g(A) (548) An example of an independent set of commuting operators is given by any of the Hermitean operators p ˆz , p ˆy or p ˆx , or alternatively ˆrz , ˆry or ˆrx . The degeneracy of simultaneous eigenvalues of a set of operators is equal to the number of eigenfunctions corresponding to the set of eigenvalues. A complete set of commuting operators is defined to be a maximal set of independent commuting Hermitean operators. Any state can be uniquely expanded in terms of the eigenfunctions of a complete set of commuting Hermitean 131

operators. What exactly constitutes a complete set of operators depends upon the physical space, i.e. whether it is one-dimensional, two-dimensional or threedimensional, and whether there is an intrinsic space associated with the particle such as spin. If the set is not complete then one can add another independent operator to the set until the set becomes complete. A state which corresponds to an eigenfunction of a complete set of commuting operators is uniquely determined and is completely defined without any remaining arbitrariness. By successively adding operators to a set of independent commuting operators, one has reduced the degeneracy until the simultaneous eigenvalues are non-degenerate and the set of operators is complete.

3.5.17

Non-Commuting Operators

It is impossible to find a set of simultaneous eigenfunctions of non-commuting operators. Thus, if ˆ ] = Cˆ [ Aˆ , B (549) and if φ(r) is a simultaneous eigenfunction then ˆ ] φ(r) = Cˆ φ(r) = 0 [ Aˆ , B

(550)

This would require that φ(r) is also an simultaneous eigenfunction of Cˆ with eigenvalue zero, which is very restrictive and very unlikely. Thus, it is impossible to know the values of A and B at the same time if they are non-commuting. An example of this is given by the momentum and position operators corresponding to the same coordinate, p ˆx and rˆx . ——————————————————————————————————

3.5.18

Exercise 42

ˆ are Hermitean operators and Show that if Aˆ and B ˆ ] = i Cˆ [ Aˆ , B

(551)

then Cˆ is Hermitean. ——————————————————————————————————

3.5.19

Solution 42

Consider the matrix elements of the commutator Z Z Z ˆ ] Ψ(r) = ˆ Ψ(r) − ˆ Aˆ Ψ(r) d3 r Φ∗ (r) Aˆ B d3 r Φ∗ (r) B d3 r Φ∗ (r) [ Aˆ , B (552) 132

On using the definition of the Hermitean conjugate of the products, we find  Z ∗  Z ∗ ˆ † Aˆ† Φ(r) ˆ † Ψ(r) = d3 r Ψ∗ (r) B − d3 r Φ∗ (r) Aˆ† B (553) ˆ are Hermitean, one has and as Aˆ and B  Z ∗ 3 ∗ ˆ ˆ = − d r Ψ (r) [ A , B ] Φ(r)

(554)

Hence, the Hermitean conjugate of the commutator is ˆ ]† = − [ Aˆ , B ˆ ] [ Aˆ , B so 

i Cˆ

†

 = −

i Cˆ

(555)

 (556)

and as the Hermitean conjugate of i is − i , we have − i Cˆ † = − i Cˆ

(557)

Thus, we recognize Cˆ as being Hermitean. ——————————————————————————————————

3.5.20

The Uncertainty Principle

The Heisenberg uncertainty relation provides a quantification of the degree to which one can determine the values of two non-commuting Hermitean operaˆ in a specific state Ψ(r). tors, Aˆ and B ˆ and form the positive definite quantity Consider the deviations ∆Aˆ and ∆B 2   Z X 3 ∗ ˆ ˆ d r Ψ (r) ∆A + i µ ∆B φn (r) ≥ 0 (558) n

where µ is a real variable, and φn (r) form a complete set of eigenfunctions. On using the completeness relation, one finds    Z ˆ ˆ Ψ(r) ∆Aˆ − i µ ∆B = d3 r Ψ∗ (r) ∆Aˆ + i µ ∆B   Z 3 ∗ 2 2 2 ˆ ˆ ˆ ˆ = d r Ψ (r) ∆A + i µ [ ∆B , ∆A ] + µ ∆B Ψ(r) (559) and since this has no zeros as a function of µ, the discriminant of this quadratic equation in µ must be greater than zero. Hence, Z Z ˆ 2 Ψ(r0 ) 4 d3 r Ψ∗ (r) ∆Aˆ2 Ψ(r) d3 r0 Ψ∗ (r0 ) ∆B  Z ≥

ˆ ] Ψ(r) d r Ψ (r) i [ ∆Aˆ , ∆B 3



133

2 (560)

Thus, we have the Heisenberg uncertainty relation Z 3 ∗ ˆ ˆ 2 ∆Arms ∆Brms ≥ d r Ψ (r) i [ A , B ] Ψ(r)

(561)

The equality only holds if the effects of the operators on the state Ψ(r) are proportional to each other. This can be seen by examining eqn(558) with the equality sign holding. This implies that, if the equality holds, the sums of the squares of the matrix elements   Z 3 ∗ ˆ ˆ d r φn (r) ∆A − i µ ∆B Ψ(r) (562)  must be zero for all n. If we expand the transformed state

ˆ ∆Aˆ − i µ ∆B

 Ψ

in terms of φn 

ˆ ∆Aˆ − i µ ∆B

 Ψ(r) =

X

Cn φn (r)

(563)

n

then we recognize that we just proved that if the equality is to hold Cn = 0 for all n. The completeness condition then implies that we have the expansion of the zero state, so   ˆ Ψ(r) = 0 (564) ∆Aˆ − i µ ∆B when the equality holds. The value of the constant, µ, can be obtained from 2 the inequality by noting that if the equality holds, then as ∆Brms > 0 the value of µ must be a repeated root. On solving the quadratic equation for the repeated root, we find R 3 ˆ ] Ψ(r) d r Ψ∗ (r) [ Aˆ , B µ = i (565) 2 2 ∆Brms Hence, on substituting the value of µ back we have the equation R 3 ˆ ] Ψ(r)  d r Ψ∗ (r) [ Aˆ , B ˆ Ψ(r) ∆Aˆ Ψ(r) = ∆B 2 2 ∆Brms

(566)

This equation is satisfied for the state Ψ(r) which produces the minimum uncertainty in ∆A and ∆B. Applying the uncertainty relation to the canonically conjugate coordinates and momenta, we have the inequality ∆pi

rms

∆rj

rms

134



¯ h δi,j 2

(567)

in which the right hand side is independent of the state of the system. The minimum uncertainty equation takes the form     ∂ i ¯h − i¯ h − px Ψ(x) = x − x Ψ(x) (568) ∂x 2 ∆x2rms which has the solution 2

  Ψ(x) =



∆x2rms

− 14

x − x

 exp



4 ∆x2rms

p x + i x ¯h

 (569)

——————————————————————————————————

3.5.21

Exercise 43

Consider a particle moving in one dimension in the presence of a potential V (x). Show that the uncertainty in the energy and position are related via ∆Erms ∆xrms =

¯h p 2m

(570)

——————————————————————————————————

3.5.22

Solution 43

The Hamiltonian is given by ˆ = H

pˆ2 + V (ˆ x) 2m

(571)

ˆ and x The commutator of H ˆ is found as ¯ h ˆ , x [H ˆ] = −i pˆ m

(572)

Then on using the generalized uncertainty principle, one has ∆Erms ∆xrms =

¯h p 2m

(573)

as was to be proved. ——————————————————————————————————

135

3.5.23

Exercise 44

Derive uncertainty relations for the various components of the orbital angular ˆ x and L ˆ y simultaneously momenta. For what states is it possible to measure L with minimum uncertainty? ——————————————————————————————————

3.5.24

Solution 44

From the commutation relation ˆx , L ˆ y ] = i ¯h L ˆz [L

(574)

one has the uncertainty relation ( ∆Lx )rms ( ∆Ly )rms ≥

¯ h Lz 2

(575)

ˆ x and L ˆ y with minimum uncertainty in Thus, it is possible to measure L ˆ z is zero states Ψ(r) where the average value of L Z ˆ z Ψ(r) = 0 Lz = d3 r Ψ∗ (r) L (576) In particular, the state with total angular momentum zero is a simultaneous ˆx, L ˆ y and L ˆ z , with all three eigenvalues being zero. eigenstate of L ——————————————————————————————————

3.5.25

Exercise 45

ˆ and Cˆ satisfy the commutation relations If the Hermitean operators Aˆ , B ˆ , Cˆ ] = i Aˆ [B ˆ [ Aˆ , Cˆ ] = i B

(577)

then show that the uncertainty relation ∆( A B )rms ∆Crms ≥

1 2



 A2 + B 2

(578)

is satisfied. ——————————————————————————————————

136

3.5.26

Solution 45

This exercise is a simple example of the uncertainty principle. The commutator ˆ with Cˆ is evaluated as of the composite operator Aˆ B ˆ , Cˆ ] = Aˆ [ B ˆ , Cˆ ] + [ Aˆ , Cˆ ] B ˆ [ Aˆ B   ˆ2 = i Aˆ2 + B

(579)

Hence, on substituting into the general statement of the uncertainty principle, one obtains the result   1 2 2 ∆( A B )rms ∆Crms ≥ A + B (580) 2 which was to be proved. ——————————————————————————————————

3.5.27

Exercise 46

ˆ z and the azimuthal Show that the z component of the angular momentum L angle ϕ satisfy the commutation relations [ [

ˆ z , cos ϕ ] = + i ¯h sin ϕ L ˆ z , sin ϕ ] = − i ¯h cos ϕ L

(581)

Hence, derive the uncertainty relations ¯ h sin ϕ 2 ¯h ≥ cos ϕ 2

∆(cos ϕ) ∆Lz ≥ ∆(sin ϕ) ∆Lz

(582)

——————————————————————————————————

3.5.28

Solution 46

The uncertainty relations applied to the azimuthal angle are problematic, since it is not clear whether the angle ϕ is really only well defined modulo 2π. An angle ϕ cannot be distinguished by a single physical measurement from ϕ + m 2 π where m is an integer. If, on the other hand, ϕ is considered to be defined only ˆz on the restricted interval (0, 2π) then the uncertainty principle for ϕ and L would read ¯h (∆ϕ)rms (∆Lz )rms > (583) 2 137

2

y

1

-2

sin φ

φ

0 -1

0

cos φ

1

2

-1

-2

x

Figure 26: The azimuthal angle ϕ of a point on a unit circle. ˆ z , which is false, or which would imply that either there are no eigenstates of L that the uncertainty of ϕ can be infinite, which is contrary to our assumption. The only way out of this quandry is to assert that an angle is not measurable, but that only the coordinates sin ϕ or cos ϕ are measurable. These functions uniquely define ϕ modulo 2 π. We need to restrict our attention to functions ˆ z is only Hermitean on the space and operators which are periodic in ϕ since L of functions which are periodic in ϕ. The commutation relations for the coordinates and angular momentum are evaluated as ˆ z , cos ϕ ] = i ¯h sin ϕ [L ˆ z , sin ϕ ] = − i ¯h cos ϕ [L

(584)

Simple substitution into the uncertainty relation leads to the equations )2

( ∆Lz )2 ( ∆ sin ϕ )2

( ∆Lz

)2

( ∆ cos ϕ





¯2 h 4 ¯2 h 4





138

2 sin ϕ 2 cos ϕ

(585)

The above equation does not produce any contradiction with the fact that   1 Φm (ϕ) = √ exp i m ϕ (586) 2π ˆ z since the expectation value of the trigonometric functions are eigenstates of L cos ϕ and sin ϕ are both zero. In this case, one finds that the uncertainty relations reduce to the equality 0 = 0. ——————————————————————————————————

139

3.6

The Philosophy of Measurement

In Quantum Mechanics the measurement process disturbs the system and is, therefore, represented by an operator. The measurement of A produces instantaneous changes from one state, say Ψ(r) of the system to another. The state of the system immediately after the measurement is that which corresponds to an eigenstate of the operator Aˆ which has the eigenvalue that corresponds to the result of the measurement a. Prior to the measurement, and the discovery of the result a, it is impossible to predict which eigenstate the system will take. All that can be predicted is the probability that the measurement will result in a particular eigenvalue a. This probability is proportional to the squared modulus of the expansion coefficient Ca X Ψ(r) = Ca φa (r) (587) a

Thus, one can view quantum mechanical measurement as a sort of filter, which instantaneously filters out the state φa (r) from the initial state of the system Ψ(r). Since this resulting state after the measurement is represented as a pure ˆ it has only one expansion coefficient. Therefore, there is a eigenfunction of A, ˆ as long probability of one for finding the result a in a further measurement of A, as this measurement is made very soon after the first measurement was made. —————————————————————————————————— Example ˆ do not commute. The operator Aˆ has The Hermitean operators Aˆ and B two non-degenerate eigenstates and eigenvalues given by Aˆ Φi (r) = ai Φi (r)

(588)

ˆ also has two non-degenerate eigenstates for i = 1 , 2. The operator B ˆ Θi (r) = bi Θi (r) B

(589)

for i = 1 , 2. The state Φ1 (r) can be expressed as a linear superposition of the eigenstates Θi (r), as Φ1 (r) = C1 Θ1 (r) + C2 Θ2 (r)

(590)

where C1 and C2 are known complex constants. (i) Find an expression for Φ2 (r) in terms of the Θi (r), and determine the expansion coefficients in terms of the known quantities Ci . A quantum mechanical particle is in a state Ψ(r) given by r r 5 1 Φ2 (r) Ψ(r) = Φ1 (r) + 6 6 140

(591)

(ii) Determine the probabilities of finding the results a1 and a2 in a meaˆ The quantity B ˆ is then measured immediately after Aˆ has been surement of A. ˆ yields the result measured. What is the probability that the measurement of B b1 ? (iii) If the measurement of Aˆ had not been performed, what is the probabilˆ will result in the measured value b1 ? ity that the measurement of B —————————————————————————————————— Solution ˆ is Hermitean the eigenstates of Aˆ can be expanded in terms of the Since B ˆ The eigenstate Φ1 (r) is expanded as eigenstates of B. Φ1 (r) = C1 Θ1 (r) + C2 Θ2 (r)

(592)

and Φ2 (r) can be expanded in terms of the complete set of Θi (r) as Φ2 (r) = D1 Θ1 (r) + D2 Θ2 (r)

(593)

where D1 and D2 are to be determined. As Φ2 (r) must be orthogonal to Φ1 (r), then Z 0 = d3 r Φ∗1 (r) Φ2 (r) Z Z = C1∗ D1 d3 r Θ∗1 (r) Θ1 (r) + C1∗ D2 d3 r Θ∗1 (r) Θ2 (r) Z Z + C2∗ D1 d3 r Θ∗2 (r) Θ1 (r) + C2∗ D2 d3 r Θ∗2 (r) Θ2 (r) (594) Since Θ1 (r) and Θ2 (r) form an orthonormal set, one has 0 = C1∗ D1 + C2∗ D2

(595)

Also, as the states are normalized to unity, one has | C1 |2 + | C2 |2 = | D1 |2 + | D2 |2 = 1

(596)

and so the unknown constants Di are found as D1 D2

= − C2∗ = C1∗

(597)

up to an arbitrary common phase. Hence, we have Φ2 (r) = − C2∗ Θ1 (r) + C1∗ Θ2 (r) 141

(598)

A measurement of Aˆ on the state r r 1 5 Ψ(r) = Φ1 (r) + Φ2 (r) 6 6 yields the result a1 with probability P (a1 ) = probability P (a2 ) = 56 .

1 6

(599)

and the result a2 occurs with

If the measurement Aˆ has been performed and the result a1 has been found, the system then is in the state Φ1 (r). The conditional probability that a subseˆ yields the result b1 is given by quent measurement of B P (a1 |b1 ) = | C1 |2

(600)

ˆ yields the result b2 and the conditional probability that the measurement of B is given by P (a1 |b2 ) = | C2 |2 (601)

However, if on the other hand the result of the first measurement is a2 , the system is in state Φ2 (r) just after the first measurement. The conditional ˆ yields the result b1 is given by probability that a subsequent measurement of B P (a2 |b1 ) = | C2 |2

(602)

ˆ yields the result b2 and the conditional probability that the measurement of B is given by P (a2 |b2 ) = | C1 |2 (603)

ˆ will result in the The total probability that the second measurement of B ˆ value b1 , no matter what value of A is measured, is given by P (A|b1 ) =

1 5 | C1 |2 + | C2 |2 6 6

(604)

ˆ is performed directly on the state Ψ, then as If the measurement of B r r r r    5 1 ∗ 1 5 ∗ C2 + C1 Θ2 (r) (605) Ψ(r) = C1 − C2 Θ1 (r) + 6 6 6 6 the result b1 occurs with probability r r 1 5 ∗ C P (b1 ) = C1 − 6 2 6 142

2

(606)

The probability P (b1 ) is not the same as P (A|b1 ) as the measurement of Aˆ disturbs the system. —————————————————————————————————— The time evolution of a system between two successive instants of time is governed by the Schr¨ odinger equation, as long as no measurements are made on the system in the time interval between the two times. In order to solve the Schr¨ odinger equation, which is a first order differential equation in the time variable, it is necessary to have one initial condition for the wave function. This is usually given by the state of the system at the initial time t0 , Ψ(r, t0 ). Once the initial condition is known, the Schr¨odinger equation will predict how the ˆ is perstate evolves with time until a measurement (say a measurement of A) formed on the system. Let us assume that the measurement occurs at time t1 . The measurement at t1 disrupts the evolution process, and sends the system into an eigenstate of the measurement operator with a definite eigenvalue. The value of the eigenvalue is the result of the measurement. The probability of getting the result a, depends on the expansion coefficient of the state Ψ(r, t1 ). After the measurement, the system has an new initial state at t1 which is φa (r). This new initial state should be used in the Schr¨odinger equation to predict the subsequent evolution of the system from t1 to later times t2 , as long as no other measurements occur in this time interval. If there is a second meaˆ the process repeats. The surement (such as a measurement represented by B), system evolves forward from the time of the second measurement according to the Schr¨ odinger equation, and the result of the second measurement provides the new initial condition at the time of the latest measurement. ——————————————————————————————————

3.6.1

Exercise 47

A particle moving in three dimensions is in a quantum mechanical state described by a wave function   2 (607) Ψ(r) = C exp − α r in the presence of a spherically symmetric potential. The Hamiltonian is given by the operator 2 2 ˆ = − ¯h ∇2 + m ω r2 H (608) 2m 2 The energy of the state is measured. Find the probability that the system is in the ground state E0 = 32 ¯h ω, where the ground state wave function φ0 (r) is given by   3  m ω r2 mω 4 exp − (609) φ0 (r) = ¯h π 2 ¯h 143

Determine the probability that the system is found to be in an excited state. What is the probability that the above system is found to be in a state φ1 (r) corresponding to E1 = 25 ¯h ω, where φ1 (r) = ( 2 π )

1 2



mω ¯h π

 54

 r cos θ exp

m ω r2 − 2 ¯h

 (610)

which is also an eigenstate of angular momentum l = 1. ——————————————————————————————————

3.6.2

Solution 47

The normalization constant for Ψ(r) is found from the equation Z 1 = d3 r | Ψ(r) |2   Z = | C |2 d3 r exp − 2 α r2 2

Z

2 π

= |C |

dϕ 0

= | C |2 4 π

π

Z

Z





2

dθ sin θ dr r exp 0   2 2 dr r exp − 2 α r

−2 α r

2



0

Z



(611)

0

Hence,  |C | =

2α π

 34 (612)

ˆ The wave function Ψ(r) can be expanded in terms of the eigenfunctions of H, which are the φn (r). Since the wave functions are normalized, the probability ˆ will result in the value E0 is given by the modulus that a measurement of H square of the expansion coefficient. The expansion coefficient for the state φ0 (r) is found from the overlap Z

3

d r

φ∗0 (r)

Ψ(r)

=

2

4π|C | 

=

4α (α +



mω ¯h π

m ω 2 h ¯ m ω 2 h ¯

 32 Z

∞ 2

dr r exp 0



mω − (α + ) r2 2 ¯h

 34 (613)

)2

The probability that the system is found in the ground state is given by  P0 =



4α (α + 144

m ω 2 h ¯ m ω 2 h ¯

 32 )2

(614)

and the probability that the system ends up in an excited state is given by Pexc

= =

1 − P0  1 −

4α (α +

m ω 2 h ¯ m ω 2 h ¯

 32 (615)

)2

Note that if α has the value

mω (616) 2 ¯h then the probability that the system will undergo a transition to a different ˆ is measured is zero. That is, the system is already in an eigenstate when H ˆ with eigenvalue E0 and any subsequent measurements of H ˆ eigenstate of H (but only if infinitesimally small times have elapsed since Ψ(r) was created) will yield the same result E0 . α =

The probability that the system ends up in the excited state φ1 (r) is zero, since on evaluating the expansion coefficient one finds Z d3 r φ∗0 (r) Ψ(r) 3 2

= (2π) |C |



mω π ¯h

 54 Z

π

Z dθ sin θ cos θ

0

∞ 3

dr r exp 0

= 0



 mω 2 − (α + )r 2 ¯h (617)

since the integral over θ vanishes identically. The vanishing of this matrix element is a natural consequence of the conservation of angular momentum for a spherically symmetric potential. The initial state is an eigenstate of the square c2 , with eigenvalue zero, and is also an eigenstate of the angular momentum, L of the z component of the angular momentum operator with eigenvalue zero. c2 The state φ1 (r) is an eigenstate of the square of the angular momentum L 2 with eigenvalue 2 ¯h and is an eigenstate of the z component of the angular momentum with angular momentum zero. Thus, these states correspond to c2 with different eigenvalues. The overlap is zero because states eigenstates of L corresponding to different eigenvalues of Hermitean operators are orthogonal. ——————————————————————————————————

3.6.3

Exercise 48

The initial state of a system is represented by the wave function  3   mω 4 m ω ( r − a )2 Ψ(r) = exp − hπ ¯ 2 ¯h

(618)

where a = eˆz a. What is the probability that after an energy measurement the system will be found in the energy eigenstates φ0 (r) and φ1 (r) given below? 145

The state φ0 (r) is given by 

mω ¯h π

φ0 (r) =

 34

 exp



m ω r2 2 ¯h

 (619)

which is a simultaneous eigenstate of energy E0 = 3/2 ¯h ω and of angular momentum with l = 0. The state φ1 (r) is given by φ1 (r) = ( 2 π )

1 2



mω ¯h π

 54

 r cos θ exp

m ω r2 − 2 ¯h

 (620)

which is an eigenstate of energy E1 = 5/2 ¯hω and is an eigenstate of angular momentum with l = 1 and m = 0. ——————————————————————————————————

3.6.4

Solution 48

The probability that the system initially in the state Ψ(r) is found in an eigenstate with energy En is found from the eigenstate expansion X Ψ(r) = Cn φn (r) (621) n

as P (En ) = | Cn |2 where

Z Cn =

d3 r φ∗n (r) Ψ(r)

(622) (623)

The expansion coefficient C0 is found as 3     a 2 a 2 mω 2 mω mω = d r exp − (r − ) exp − (r + ) π ¯h 2 ¯h 2 2 ¯h 2   mω 2 = exp − a (624) 4 ¯h Z

C0

3



Thus, we find that the probability of finding the result E0 if the energy of state Ψ(r) is measured is given by   mω 2 P (E0 )l=0 = exp − (625) a 2 ¯h which decreases exponentially with increasing a.

146

The expansion coefficient C1 is determined from C1

2   mω az mω a2 2 = (2π) d r (z + ) exp − (r + ) π ¯h 2 ¯h 4  1   m ω a2z 2 mω 2 = exp − a (626) 2 ¯h 4 ¯h 1 2

Z

3



where az is the z component of the vector displacement a. Hence, one has the probability of finding a simultaneous eigenstate of energy E1 and angular momentum l = 1 and m = 0 is given by     mω 2 m ω a2z P (E1 )l=1,m=0 = exp − a (627) 2 ¯h 2 ¯h

——————————————————————————————————

3.6.5

Exercise 49

A particle is confined to move in a one-dimensional interval between x = 0 and x = L. The initial state is given by r 30 Ψ(x) = x(x − L) (628) L5 for L > x > 0, and 0 otherwise. Find the probability that, after a measurement of the energy, the system will be found in the eigenstate r 2 nπx φn (x) = sin (629) L L corresponding to energy En =

h ¯ 2 π 2 n2 2 m L2 2

2

If the energy is measured and found to be equal 9 2 h¯mπL2 , what is the probability density that in a second measurement the particle will then be found in an interval of width δx around x = L2 ? If the measurement of En had not taken place, what would the probability density for finding the particle at x = L2 have been? ——————————————————————————————————

147

3.6.6

Solution 49

First we shall evaluate the indefinite integrals Z x Z x ∂ dx0 x0 sin α x0 = − dx0 cos α x0 ∂α sin α x − α x cos α x = (630) α2 and Z x Z x ∂2 dx0 x02 sin α x0 = − dx0 sin α x0 ∂α2 ( 2 − α2 x2 ) cos α x + 2 α x sin α x = α3 (631) Then the coefficients Cn in the expansion of Ψ(x) in terms of the eigenfunctions φn (x) are given by Z L Cn = dx0 φ∗n (x) Ψ(x) √0 Z L nπx 60 dx0 x ( x − L ) sin = L3 L 0 √ 60 2 ( cos n π − 1 ) (632) = n3 π 3 Hence, only terms with odd values of n appear in the expansion. The probability of observing the eigenvalue En is only finite for odd n and is given by P (n) P (n)

= | Cn2 | 480 = ( 1 − cos n π ) n6 π 6

(633)

Since the energy measurement picks out the non-degenerate eigenfunction corresponding to n = 3, the probability density for finding the particle at x = L2 is given by 2 P (L/2) = (634) L whereas if the measurement had not occurred then we would have anticipated that the probability density would be 15 P (L/2) = (635) 8L Thus, the measurement of the energy has changed the probabilities of the measurement of the particle’s position made at an instant later. ——————————————————————————————————

148

3.7

Time Evolution

The evolution equation of a physical quantity B(p, q, t) in Classical Mechanics is given in terms of the Poisson Brackets as, dB ∂B = [ B , H ]P B + dt ∂t

(636)

where H is the Hamiltonian, which is a function of the canonically conjugate momenta p, coordinates q and possibly t. We regard this as an equation of motion for the expectation value. To obtain the quantum mechanical equation of motion from the classical equation, we replace the classical quantity B(p, q, t) ˆ = B(ˆ by the operator B p, qˆ, t), and the classical Hamiltonian H(p, q, t) by the ˆ Hamiltonian operator H = H(ˆ p, qˆ, t). In the non-relativistic limit, the classical Hamiltonian for a particle, of charge q, in an electromagnetic field is given by  2 1 q H = p − A(r, t) + q φ(r, t) (637) 2m c where we are now representing the position vector as r. The Hamiltonian operator is  2 q 1 ˆ pˆ − A(ˆ r, t) + q φ(ˆ r, t) (638) H = 2m c which in the coordinate representation becomes 2  1 q ˆ H(r, t) = − i ¯h ∇ − A(r, t) + q φ(r, t) 2m c

(639)

On substituting operators for the classical variables in the classical equation of motion, replacing the Poisson Bracket with the commutator divided by the i ¯h, and taking the expectation value in a state Ψ(r, t), one has Z Z d ˆ t) Ψ(r, t) = ˆ t) , H(r, ˆ t) ] Ψ(r, t) d3 r Ψ∗ (r, t) B(r, d3 r Ψ∗ (r, t) [ B(r, i ¯h dt Z ∂ ˆ B(r, t) Ψ(r, t) + i ¯h d3 r Ψ∗ (r, t) ∂t (640) This should be regarded as an equation which governs the time dependence of the expectation value of B(t) Z ˆ t) Ψ(r, t) B(t) = d3 r Ψ∗ (r, t) B(r, (641) The time dependence of the expectation value B(t) has a formal solution given by Z ˆ † (t, 0) B(r, ˆ t) U ˆ (t, 0) Ψ(r, 0) B(t) = d3 r Ψ∗ (r, 0) U (642) 149

ˆ (t, t0 ) is the time evolution operator which describes the implicit time where U dependence of the system and satisfies the first order equation i¯ h

d ˆ ˆ t) U ˆ (t, t0 ) U (t, t0 ) = H(r, dt

(643)

ˆ (t0 , t0 ) = 1 and U ˆ † (t, t0 ) is its Hermitean conjuwith the boundary condition U gate. From the solution given in eqn(642) one can derive a number of “pictures” of quantum mechanics. The two most popular are : (i) The Schr¨ odinger picture. (ii) The Heisenberg picture.

3.7.1

The Schr¨ odinger Picture.

In this picture, the operators are independent of time (the only exception is if there is some externally imposed time dependence, like a time-dependent electromagnetic field etc.). That is, the operators only change through their explicit time dependence. The implicit time dependence of the system is carried in the wave function. Thus, in the Schr¨odinger picture ˆ (t, t0 ) Ψ(r, t0 ) Ψ(r, t) = U ˆ t) = B(r, ˆ t0 ) B(r,

(644)

The time evolution is usually referenced with respect to an initial time t0 which ˆ (t, t0 ) is unitary, is usually chosen as t0 = 0. If the time evolution operator U the wave function will retain its initial normalization. From this, one finds that the implicit time dependence is contained in the wave function and, is governed by the first order differential equation i¯ h

d ˆ t) Ψ(r, t) Ψ(r, t) = H(r, dt

(645)

with one initial condition. In the momentum space representation, the Schr¨odinger equation has the form  2  p d i¯ h Φ(p, t) = + q φ( i ¯h ∇p , t) Φ(p, t) (646) dt 2m

——————————————————————————————————

150

3.7.2

Exercise 50

Derive the momentum space form the Schr¨odinger equation from the real space form. ——————————————————————————————————

3.7.3

Solution 50

The time-dependent Schr¨ odinger equation is i¯ h

¯h2 ∂Ψ(r, t) = − ∇2 Ψ(r, t) + q φ(r, t) Ψ(r, t) ∂t 2m

(647)

in position representation, where Ψ(r, t) is the time-dependent wave function. The time-dependent momentum space wave function Φ(p, t) is related to Ψ(r, t) via the generalized Fourier transform    32 Z  p0 . r 1 d3 p0 Φ(p0 , t) exp + i (648) Ψ(r, t) = 2π¯ h ¯h On substituting this relation in the first term on the right hand side, one obtains the contribution   32 Z   p02 p0 . r h2 ¯ h2 ¯ 1 − ∇2 Ψ(r, t) = d3 p0 Φ(p0 , t) 2 exp + i 2m 2m 2 π ¯h ¯h ¯h   32 Z   02 0 p p .r 1 = d3 p0 Φ(p0 , t) exp + i 2 π ¯h 2m ¯h (649)

On performing the inverse Fourier Transform   32 Z   p.r 1 3 d r Ψ(r, t) exp − i Φ(p, t) = 2 π ¯h ¯ h

(650)

on the time-dependent Schr¨odinger equation, one has  3 Z   02 Z ∂Φ(p, t) ( p0 − p ) . r p 1 3 3 0 i ¯h = d r d p exp i Φ(p0 , t) ∂t 2π¯ h 2m ¯h    32 Z  p.r 1 3 +q d r Ψ(r, t) φ(r, t) exp − i 2 π ¯h ¯h (651) where we have switched the order of the wave function and the scalar potential in the last term. On recognizing the integral over r as being proportional to the 151

integral representation of the delta function δ 3 (p − p0 ), and then performing the integration over p, one obtains i¯ h

∂Φ(p, t) ∂t

=

p2 Φ(p, t) 2m   32 Z   p.r 1 +q d3 r Ψ(r, t) φ(r, t) exp − i 2 π ¯h ¯ h (652)

The last term can be manipulated as p2 Φ(p, t) 2m    32 Z  p.r 1 3 +q d r Ψ(r, t) φ(i ¯h ∇p , t) exp − i 2 π ¯h ¯h 2 p Φ(p, t) = 2m    32 Z  p.r 1 + q φ(i ¯ h ∇p , t) d3 r Ψ(r, t) exp − i ¯h 2 π ¯h 2 p = Φ(p, t) + q φ(i ¯h ∇p , t) Φ(p, t) (653) 2m In the last line we have used the definition of the inverse Fourier transform. Thus, we have derived the equation that governs the time dependence of the momentum space wave function   2 ∂Φ(p, t) p i¯ h = + q φ(i ¯h ∇p , t) Φ(p, t) (654) ∂t 2m

i ¯h

∂Φ(p, t) ∂t

=

analogous to the Schr¨ odinger equation in the position representation. ——————————————————————————————————

3.7.4

The Heisenberg Picture.

In the Heisenberg picture, the wave function is chosen to be independent of time and the operators carry the implicit time dependence which represents the dynamics of the system. Thus, in this picture we have Ψ(r, t) = Ψ(r, t0 ) = Ψ(r, 0) ˆ † (t, t0 ) B(r, ˆ t) U ˆ (t, t0 ) ˆ B(r(t), t) = U

(655)

Hence, the equation of motion for the implicit time dependence comes from the equation of motions for the operators and is given by i¯ h

d ˆ ∂ ˆ ˆ ˆ t) , H(r(t), t) ] + i h ¯ B(r(t), t) B(r(t), t) = [ B(r(t), ∂t dt 152

(656)

——————————————————————————————————

3.7.5

Exercise 51

Find the equation of motion for the position operator rˆ(t) in the Heisenberg picture for a system with the Hamiltonian given by ˆ H(t) =

p ˆ(t)2 + V (ˆ r(t)) 2m

(657)

——————————————————————————————————

3.7.6

Solution 51

We shall examine the Heisenberg equation of motion i¯ h

dˆ r ∂ rˆ ˆ = [ rˆ(t) , H(t) ] + ih ¯ dt ∂t

(658)

at the instant of time, say t = 0. Then in the position representation with rˆ(0) = r and pˆ(0) = − i ¯h ∇ one has no explicit time dependence for rˆ and the implicit dependence is governed by i¯ h

dr dt

ˆ ] = [r, H p ˆ2 ] + [ r , V (r) ] 2m pˆ = 2ih ¯ 2m

=

[r,

Thus, we find the operator equation pˆ dˆ r = dt m

(659)

This operator equation has the same form as the momentum velocity relation in classical mechanics. ——————————————————————————————————

153

3.7.7

Exercise 52

Find the Heisenberg picture equation of motion for the momentum operator pˆ(t) for a system with the Hamiltonian given by ˆ = H

p ˆ2 + V (ˆ r) 2m

(660)

——————————————————————————————————

3.7.8

Solution 52

We shall examine the Heisenberg equation of motion i ¯h

dˆ p ∂ pˆ ˆ = [ pˆ(t) , H(t) ] + ih ¯ dt ∂t

(661)

at the instant of time, say t = 0. Then in the position representation with rˆ(0) = r and pˆ(0) = − i ¯h ∇, one has no explicit time dependence for pˆ and the implicit dependence is governed by i¯ h

dˆ p dt

=

ˆ ] [ pˆ , H

p ˆ2 ] + [ pˆ , V (r) ] 2m = − i ¯h ∇ V (r) = [ pˆ ,

(662)

Thus, we find the operator equation dˆ p = − ∇ V (r) dt

(663)

This equation has the same form as Newton’s laws, except the classical quantities are replaced by operators. —————————————————————————————————— The rate of change with respect to time of the operators found in Exercises 51 and 52 should look exactly like Hamilton’s equations of motion in classical mechanics. ——————————————————————————————————

154

3.7.9

Exercise 53

Prove that in the Heisenberg picture the rate of change of the angular momentum operator satisfies the equation d ˆ L(t) = − q r(t) ∧ ∇φ(r(t), t) dt

(664)

which is similar to the corresponding equation involving the torque classical mechanics. ˆ does not explicitly depend on time t, then the Heisenberg If an operator B equation of motion reduces to i¯ h

d ˆ ˆ ˆ B(r(t)) = [ B(r(t)) , H(r(t), t) ] dt

(665)

ˆ commutes with and so we find that the operator does not change with time if B ˆ the Hamiltonian H. As the wave function is time independent in the Heisenberg picture, the expectation values of the operator will also be time independent. Thus, B will be a constant of motion, or a conserved quantity, if ˆ , H ˆ ] = 0 [B

(666)

ˆ for B, ˆ one Since any operator commutes with itself, when one substitutes H finds that the Hamiltonian commutes with itself. Thus, if the Hamiltonian has no explicit dependence on time then the energy is a constant of motion. ——————————————————————————————————

3.7.10

Exercise 54

ˆ which has no explicit time Find the time dependence of a Hermitean operator B dependence, if the commutator with the Hamiltonian is a complex constant C, i.e. ˆ , B ˆ ] = C [H (667)

——————————————————————————————————

3.7.11

Solution 54

ˆ is governed by the Heisenberg equation The time dependence of the operator B of motion ˆ dB ˆ , H ˆ ] i ¯h = [B (668) dt 155

Hence, on using the commutation relation i ¯h which leads to

ˆ dB = −C dt

C ˆ ˆ B(t) = B(0) + i t ¯h

(669)

(670)

This has an application for the rate of change of momentum p ˆ for a particle in a uniform applied field F , ˆ = H

p ˆ2 − F .r 2m

(671)

which leads to the commutation relation ˆ ] = + i ¯h F [p ˆ, H

(672)

Thus, we have the time dependence of the momentum operator p ˆ(t) = p ˆ(0) + F t

(673)

which is analogous to the classical solution for the momentum of a particle in an applied uniform field. ——————————————————————————————————

3.7.12

Exercise 55

Consider a free particle moving in one dimension, calculate the r.m.s. position ∆xrms (t) as a function of time, by using the Heisenberg equations of motion repeatedly. Do not assume any particular form for the wave packet. Show that ( ∆xrms (t) )2

( ∆prms (t) )2

( ∆xrms (0) )2 +   2 1 + x pˆ + pˆ x (0) − x (0) pˆ (0) t + m 2 ( ∆prms (0) )2 2 + t m2 = ( ∆prms (0) )2 (674) =

The average value of the momentum and the r.m.s. momenta are time independent as a consequence of p being conserved. ——————————————————————————————————

156

3.7.13

Exercise 56

Consider a one-dimensional harmonic oscillator with the Hamiltonian. ˆ = H

pˆ2 m ω2 2 + x 2m 2

(675)

Derive the equations of motion for the expectation value of x. Solve this equation to show that the mean position oscillates p(0) sin ω t mω

x(t) = x(0) cos ω t +

(676)

and that the r.m.s. position is given by 2

2

∆x (t)

∆p (0) 2 = ∆x (0) cos2 ω t + sin2 ω t + 2 ω2 m   1 sin 2 ω t + x pˆ + pˆ x (0) − x (0) pˆ (0) 2 mω

(677)

Show that this result reduces to that of the previous exercise when ω → 0. —————————————————————————————————— The Schr¨ odinger and Heisenberg pictures are equivalent since the only physically important quantities are measurable and, therefore, are in the form of an expectation value, as displayed in eqn(642). We shall mainly be concerned with the Schr¨ odinger picture, in the position space representation.

3.7.14

The Schr¨ odinger Equation

In the Schr¨ odinger picture, the operators are time independent and the wave functions evolve with time according to ˆ (t, t0 ) Ψ(r, t0 ) Ψ(r, t) = U

(678)

The time-dependent wave functions satisfy the Schr¨odinger equation, ∂ Ψ(r, t) ∂t ∂ ih ¯ Ψ(r, t) ∂t i¯ h

ˆ t) Ψ(r, t) = H(r, 2    q 1 + q φ(r, t) Ψ(r, t) A(r, t) = − i ¯h ∇ − c 2m (679)

This linear partial differential equation is first order in time and, therefore, requires one initial condition. At the initial time t0 = 0 the initial condition is given by knowledge of Ψ(r, 0). For the case where the Hamiltonian contains a

157

time-independent scalar electrostatic potential φ(r), the Schr¨odinger equation has the form   ∂ ¯h2 2 i¯ h Ψ(r, t) = − ∇ + q φ(r) Ψ(r, t) (680) ∂t 2m ˆ is time independent, one has a formal In this case, where the Hamiltonian H solution Ψ(r, t)

ˆ (t, 0) Ψ(r, 0) = U  ˆ t  H = exp − i Ψ(r, 0) ¯h

(681)

which contains the wave function at the initial time. It should be noted that, ˆ (t, 0) satisfies the equation since the time evolution operator U ˆ † (t, 0) U ˆ (t, 0) = Iˆ U

(682)

it is a unitary operator. The unitarity condition ensures that the normalization ˆ (t, 0) Ψ(r, 0) is independent of time. of the wave function Ψ(r, t) = U Let us now consider a Hamiltonian which contains a time-dependent external field, due to some classical source. For example, pˆ2 + V (ˆr) − ( F . rˆ ) sin ω t 2m

ˆ H(t) =

(683)

We should be aware that the Hamiltonian at different times does not commute   ˆ 1 ) , H(t ˆ 2) ] = − i ¯ [ H(t h ( F . pˆ ) sin ω t1 − sin ω t2 (684)

ˆ When the Hamiltonian has an explicit time dependence, H(t), due to an external field then the solution of the Schr¨odinger equation can be expressed as ˆ (t, 0) Ψ(r, 0) Ψ(r, t) = U

(685)

where the time dependence of the evolution operator is governed by the equation i¯ h

∂ ˆ ˆ ˆ (t, 0) U (t, 0) = H(t) U ∂t

On integrating the time evolution equation, one obtains Z t ˆ (t, 0) = 1 − i ˆ 1) U ˆ (t1 , 0) U dt1 H(t ¯h 0 158

(686)

(687)

ˆ on both sides of the equation. We This involves the time evolution operator U ˆ (t, 0) with the shall iterate this equation, that is, substitute the expression for U ˆ (t1 , 0) in the right hand two terms on the right hand side as an expression for U side. After the substitution, the equation now possesses two completely known ˆ and the second linear in H ˆ and the unknown term terms, one of order zero in H ˆ appears in a third term and has a coefficient which is proportional to involving U ˆ 2 . On further iterating this equation an infinite number of times, one obtains H ˆ an infinite series in the operator H, Z t Z t Z t0 1 i ˆ ˆ ˆ 1 ) H(t ˆ 2 ) + . . . (688) dt1 H(t1 ) − 2 dt1 dt2 H(t U (t, 0) = 1 − h 0 ¯ ¯h 0 0 where one recognizes that the Hamiltonian at earlier times always appears to the right of Hamiltonian at later times. The general higher order term in this infinite series is given by  n Z t Z t1 Z tn−1 i ˆ 1 ) H(t ˆ 2 ) . . . H(t ˆ n ) (689) + − dt1 dt2 . . . dtn H(t h ¯ 0 0 0 which we again note has a time ordered structure, in that the time tm+1 is ˆ m+1 ) is on the right of H(t ˆ m ). In the always earlier than the time tm , and H(t general term in this expansion, the intermediate times in the integration are always ordered and the Hamiltonian operators are also time ordered. We could permute the dummy variables t1 to tn in many ways, in fact there are n! such permutations. Each permutation would result in a different time ordering of the labels, and each permutation gives the same result as long as the Hamiltonians are ordered such that the Hamiltonians at consecutive times are adjacent to each other with the earlier time to the right. This procedure is called time ordering and was first introduced by the Gian-Carlo Wick. We shall ˆ which time orders a string of Hamiltonians according to define an operator W the time label. We note that each of the n! permutations involve integrations over different sectors of the “space” defined by the set of times (t1 , t2 , . . . , tn ) where t > tm > 0, and that the n! sectors completely run over the entire “volume”, tn , of this space. Hence, we can re-write the general term as an integration over the entire volume, as long as we keep the Hamiltonian in the ˆ, correct time ordered places, and this is done by following the instruction W  n Z t Z t Z t 1 i ˆ H(t ˆ 1 ) H(t ˆ 2 ) . . . H(t ˆ n) dt1 dt2 . . . dtn W + − n! h ¯ 0 0 0 (690) On summing the series, we find ˆ (t, 0) = W ˆ exp U



i − ¯h

Z

t

ˆ 0) dt H(t 0

 (691)

0

This is the best we can do for an arbitrary time-dependent Hamiltonian. Howˆ has no explicit time dependence, then the effect of W ˆ can be neglected ever, if H 159

ˆ m ) are the same. In this case we recover our previous result, since all the H(t namely   Z t ˆ (t, 0) = exp − i ˆ U dt0 H ¯h 0  ˆ t  H = exp − i (692) ¯h

——————————————————————————————————

3.7.15

Exercise 57

Find the time dependence of a state representing a particle in a field free environment, with the initial condition that at t = 0 the particle is in momentum eigenstate given by  Ψ(r, 0) =

1 2 π ¯h

 32

p0 . r

 exp

+ i

 (693)

¯h

——————————————————————————————————

3.7.16

Solution 57

The time dependence of the wave function is given by the Schr¨odinger equation i ¯h

d ˆ Ψ(r, t) Ψ(r, t) = H dt

(694)

The solution can found first by integrating Ψ(r, t) = Ψ(r, 0) −

i ¯h

Z

t

ˆ Ψ(r, t0 ) dt0 H

(695)

0

and then by iterating Ψ(r, t) = Ψ(r, 0) −

i h ¯

Z 0

t

ˆ Ψ(r, 0) − 1 dt0 H ¯h2

Z 0

t

dt0

Z

t0

ˆ 2 Ψ(r, t”) dt” H

0

(696) The equation simplifies, since the initial wave function Ψ(r, 0) is also an eigenˆ state of the free-particle Hamiltonian H ˆ = H

p ˆ2 2m

160

(697)

so that ˆ Ψ(r, 0) = E(p ) Ψ(r, 0) H 0 where

(698)

p20

E(p0 ) =

(699)

2m

ˆ in eqn(696) and is the energy eigenvalue. On substituting the eigenvalue for H iterating an infinite number of times, one finds that   E(p0 ) t Ψ(r, 0) Ψ(r, t) = exp − i ¯h   32    p0 . r − E(p0 ) t 1 exp + i (700) = 2 π ¯h ¯h

——————————————————————————————————

3.7.17

Time Development of a Wave Packet

A wave packet Ψ(r, t) representing a free particle evolves according to i¯ h

∂ ¯h2 Ψ(r, t) = − ∇2 Ψ(r, t) ∂t 2m

(701)

since the potential is a constant q φ(r, t) = V0 which can be absorbed into the reference energy, and can be set to zero. The evolution equation can be solved by Fourier decomposition, since for free particles the momentum eigenfunctions are also eigenfunctions of the Hamiltonian. By Fourier decomposition of the initial wave function Ψ(r, 0), one finds Φ(p, 0) as the coefficient of the momentum eigenfunction, φp (r). Then  φp (r) ∝ exp

i

p.r ¯h

 (702)

the wave function in momentum representation evolves with time according to   Et Φ(p, t) = exp − i Φ(p, 0) (703) ¯h p2

where E = 2 m . Thus, as the initial value of the wave function in momentum representation is given by  Φ(p, 0) =

1 2π¯ h

 32 Z

d3 r0 Ψ(r0 , 0) exp

161

 − i

p . r0 ¯h

 (704)

we find that the time development of the wave function in momentum representation is governed by      32 Z  p . r0 + E t 1 (705) Φ(p, t) = d3 r0 Ψ(r0 , 0) exp − i ¯h 2π¯ h On performing the inverse Fourier transform of Φ(p, t), we obtain the wave function in the position representation at time t as, Ψ(r, t) where  Ψ(r, t) =

1 2π¯ h

 23 Z

d3 p Φ(p, 0) exp



 + i

p.r −

p2 t 2m



 / ¯h (706)

Thus, all one needs to do is to evaluate the Fourier transform Φ(p, 0) of the initial condition, eqn(704), and the inverse transform, given in eqn(706), to obtain the time dependence of the free particle wave function. ——————————————————————————————————

3.7.18

Exercise 58

Given the initial wave function   34     p0 . r 1 ( r − r0 )2 exp − exp + i (707) Ψ(r, 0) = 2 π δx2 4 δx2 ¯ h find the time evolution of the wave function for a free particle, the time dependence of the average position, r(t), and the mean squared deviation of the particle’s position, ∆r(t)2 . ——————————————————————————————————

3.7.19

Solution 58

The momentum space wave function, at time t = 0 is found from the initial wave function in real space via the Fourier transform Φ(p, 0)



1 2π¯ h

 32 Z



2 δx2 π¯ h2

 34

= =

 p.r d r Ψ(r, 0) exp − i ¯h    2 2  ( p − p0 ) δx r 0 . ( p − p0 ) exp − exp − i ¯ h ¯h2 (708) 

3

162

Dispersion of a Gaussian wave packet 1.5

l Ψ(x,t) l2

t=0

t = t0

1

t = 2t0

t = 3t0

t = 4t0

0.5

0 -1.5

-1

-0.5

0

x

0.5

1

1.5

Figure 27: The time dependence of the probability density for an initial Gaussian p wavepacket, Ψ(x, t). The wavepacket moves with velocity m and disperses with increasing t. Then, the momentum space wave function at time t is given by 3   ( p − p0 )2 δx2 2 δx2 4 exp − π¯ h2 ¯2 h     r 0 . ( p − p0 ) p2 t × exp − i exp − i ¯h 2 m ¯h  Φ(p, t)

=

(709) The time dependence only enters via the phase. Since the momentum probability distribution involves the modulus squared wave function | Φ(p, t) |2 , we find that the momentum distribution is time independent. Furthermore the average momentum is p0 and the distribution has a r.m.s. width, in each of the three orthogonal directions, which is equal to 2 h¯δx , independent of time. The wave packet does not spread in momentum space. The real space wave function at time t is given by the inverse Fourier transform of Φ(p, t)  Ψ(r, t) =

1 2π¯ h

 32 Z

3

d p Φ(p, t) exp

163



p.r + i ¯h

 (710)

This has the explicit form   32 Z   3  ( p − p0 )2 δx2 1 2 δx2 4 3 Ψ(r, t) = d p exp − 2π¯ h π ¯h2 ¯h2       r 0 . ( p − p0 ) p2 t p.r exp − i exp + i × exp − i ¯h 2 m ¯h ¯ h (711) Combining the exponential factors and completing the square by changing variable from p to z where p

z = p − p0 one finds that  Ψ(r, t) =

1 2 π ( δx2 + i t

i ( r − r0 − t m0 ) − 2 ( δx2 + i t 2 h¯m )  34

h ¯ 2 m

( r − r0 −

 exp

)



1 m

4 ( δx2 + i t

 × exp

(712)

+ i

( p0 . r −

p0 t )2 h ¯ 2 m ) p2 2 m t



)



¯h (713)

From this we see that the wave packet corresponds to a real space probability distribution function in which the maximum moves along the classical trajectory, p0 t

(714) m The width of the distribution function, however, increases with increasing t, corresponding to a spreading of the wave packet. The root mean squared width is given by s r = r0 +

¯2 h (715) 4 m2 which increases linearly with t for sufficiently large t. This corresponds to the spread in the positions of a set of classical particles which have a momentum distribution of width 2 h¯δx . ∆x2 (t)rms =

δx4 + t2

——————————————————————————————————

3.7.20

Time Evolution and Energy Eigenfunctions

The analysis of the time development of a wave packet for a free particle can be extended to motion in an arbitrary potential, with the knowledge of the energy eigenfunctions of the time-independent Hamiltonian operator. ˆ φn (r) = En φn (r) H 164

(716)

The energy eigenstates evolve according to the Schr¨odinger equation i¯ h

∂ ˆ φn (r, t) φn (r, t) = H ∂t = En φn (r, t)

(717)

since they are eigenstates. The solution of the Schr¨odinger equation yields the time dependence of the energy eigenstates as   En t φn (r, t) = exp − i φn (r) (718) ¯h The energy eigenstates are often called stationary states, as the only time dependence occurs through a phase factor, which is generally unobservable. Given that the eigenstates of the Hamiltonian form a complete set, we can expand the initial wave function in terms of the energy eigenstates X Ψ(r, 0) = Cn φn (r) (719) n

where the expansion coefficients are given by Z Cn = d3 r0 φ∗n (r0 ) Ψ(r0 , 0)

(720)

Hence, we have the formal solution Ψ(r, t)

=

X



En t − i ¯h

Cn φn (r) exp

n

Ψ(r, t)

=

X Z

3 0

d r φn (r, 0)

φ∗n (r0 )

n

 

exp

En t − i ¯h



Ψ(r0 , 0) (721)

This expression for the time-dependent wave function involves an infinite sum of energy eigenfunctions, and satisfies the time-dependent Schr¨odinger equation as can be seen by direct substitution. We are implicitly assuming that the summation converges for all times, including t = 0 where the sum reduces to the Dirac delta function δ( r0 − r ) (722) due to the completeness relation. Thus, at t = 0, the integration over r0 yields the initial wave function Ψ(r, 0). ——————————————————————————————————

165

3.7.21

Exercise 59

A two-level system has energy eigenfunctions φ1 and φ2 with energy eigenvalues E1 and E2 respectively. The operator Aˆ does not commute with the Hamiltonian and has eigenfunctions θ1 and θ2 corresponding to the eigenvalues a1 and a2 . If the system is in the eigenstate θ1 at time t = 0 (i.e. Ψ(r, 0) = θ1 (r)), show that it is possible for the expectation value of Aˆ to obey the equation     a1 + a2 a1 − a2 ( E1 − E2 ) t a(t) = + cos (723) 2 2 ¯h where the expectation value is defined by Z a(t) = d3 r Ψ∗ (r, t) Aˆ Ψ(r, t)

(724)

Also find the probabilities that a measurement of Aˆ at time t will give the result a1 or a2 , respectively. ——————————————————————————————————

3.7.22

Solution 59

The principle of linear superposition, together with the Schr¨odinger equation for the time-independent Hamiltonian yields, the wave function of the system as     E1 t E2 t Ψ(r, t) = C1 φ1 (r) exp − i + C2 φ2 (r) exp − i ¯h ¯h (725) At t = 0 the system is in an eigenstate of Aˆ with eigenvalue a1 . Thus, θ1 (r) = C1 φ1 (r) + C2 φ2 (r)

(726)

By orthogonality of the eigenstates of Aˆ we have θ2 (r) = C2∗ φ1 (r) − C1∗ φ2 (r)

(727)

ˆ we Hence, on solving for the energy eigenstates in terms of the eigenstates of A, find that the state of the system is given by      E1 t E2 t | C1 |2 exp − i + | C2 |2 exp − i θ1 (r) Ψ(r, t) = ¯h ¯h      E1 t E2 t + C1 C2 exp − i − exp − i θ2 (r) ¯h ¯h (728) 166

The expectation value of Aˆ in this state is given by Z a(t) = d3 r Ψ∗ (r, t) Aˆ Ψ(r, t)   ( E1 − E2 ) t 4 4 2 2 = a1 | C1 | + | C2 | + 2 | C1 | | C2 | cos ¯h   ( E1 − E2 ) t + a2 2 | C1 |2 | C2 |2 1 − cos (729) ¯h This yields the required result if | C1 |4 + | C2 |4 =

1 2

(730)

and

1 (731) 2 On using the normalization condition in the last equation, one finds a quadratic equation. The solution of the quadratic equation leads to 2 | C1 |2 | C2 |2 =

| C1 |2 = | C2 |2 =

1 2

(732)

The probability of finding the result a1 is given by   ( E1 − E2 ) t 4 4 2 2 P (a1 ) = | C1 | + | C2 | + 2 | C1 | | C2 | cos ¯h (733) and the probability of finding the result a2 is   ( E1 − E 2 ) t P (a2 ) = 2 | C1 |2 | C2 |2 1 − cos ¯h (734) Note that, due to the normalization, the sum of the probabilities is unity  P (a1 ) + P (a2 ) =

| C1 |2 + | C2 |2

2 = 1

(735)

and independent of t. ——————————————————————————————————

167

3.7.23

The Correspondence Principle

The correspondence principle, as previously stated, implies that quantum mechanics should reduce to classical mechanics in the limit h ¯ → 0. We shall show how quantum mechanics reproduces the Hamilton-Jacobi equations of classical mechanics, in this limit. Starting from the Schr¨ odinger equation in the case where the vector potential is absent   ¯h2 ∂ Ψ(r, t) = − ∇2 + q φ(r, t) Ψ(r, t) (736) i¯ h ∂t 2m one can write the complex wave function in terms of two real functions A(r, t) and S(r, t) representing the amplitude and phase   i S(r, t) (737) Ψ(r, t) = A(r, t) exp + ¯h Then, on substituting this form in the Schr¨odinger equation and taking the real parts and the imaginary parts, one has two equations. The first represents the real part of the Schr¨ odinger equation  −

∂S ∂t

 =

1 2m

2

 ∇ S(r, t)

¯ 2 ∇2 A h 2m A

(738)

   ∇A . ∇S

(739)

+ q φ(r, t) −

and the second represents the imaginary part ¯h ¯h ∂A = −i A ∇2 S − ih ¯ ∂t 2m m



On taking the limit h ¯ → 0, one finds that the equation for the phase S satisfies the Hamilton - Jacobi equation of classical mechanics,  −

∂S ∂t

 =

1 2m

2

 ∇ S(r, t)

+ q φ(r, t)

(740)

In the Hamilton - Jacobi formulation of classical mechanics one identifies the classical momentum with the gradient of S p = ∇S The equation for the amplitude A can be reduced to  2   2 A ∇S ∂A = −∇. m ∂t

(741)

(742)

by multiplying by an integrating factor of A(r, t). This last equation is identified as a continuity equation which relates the probability density ρ(r, t) = 168

| Ψ(r, t) |2 = A(r, t)2 , to the probability current density j(r, t) defined by ∇ S j(r, t) = m ρ. The continuity equation is ∂ ρ(r, t) + ∇ . j(r, t) = 0 ∂t

(743)

Thus, to lowest order in h ¯ , the phase of the wave function satisfies the Hamilton - Jacobi equation and the amplitude satisfies a continuity equation. Applying this analysis to a wave packet, one finds that the wave packet follows a trajectory which is governed by the classical action S that appears in the exponent. The phase factor has the slowest variation along the classical trajectory. The continuity condition shows that the probability density ρ will vanish outside the region that moves along the classical trajectory. Furthermore, the value of any physical quantity can be approximated by the classical value if the wave function is dominated by the exponential factor. In this case one can consider the unitary transformation, which when applied to the momentum operator results in     iS iS exp − p ˆ exp + = p ˆ + ∇S (744) h ¯ ¯h This leads to the operator being represented by the classical value of the momenˆ = B(r, pˆ) tum and a quantum correction. Thus, for an arbitrary operator B one has     iS iS exp − B(r, p ˆ) exp + = B(r, p ˆ + ∇S) (745) ¯h ¯h If the amplitude A(r, t) has a sufficiently slow variation, then the expectation ˆ can be replaced by its classical value, since value of B     Z Z iS iS ˆ Ψ(r) = d3 r Ψ∗ (r) B d3 r A exp − B(r, p ˆ) exp + A ¯h ¯h Z = d3 r A B(r, p ˆ + ∇S) A ≈ B(r, ∇S)

(746)

where the real function B is evaluated on the classical trajectory on which A is non-vanishing. Let us examine the conservation of probability density that is inherent in the exact quantum mechanical Schr¨odinger equation.

3.7.24

The Continuity Equation and Particle Conservation

An exact continuity equation can be derived from the Schr¨odinger equation, which ensures that the number of particles (i.e. the probability of finding the 169

particle anywhere), is conserved. The form of the equation is ∂ ρ(r, t) + ∇ . j(r, t) = 0 ∂t

(747)

where ρ(r, t) = Ψ∗ (r, t) Ψ(r, t) is the probability density and (in the absence of a vector potential) the probability current density j(r, t) is given by j(r, t) =

¯ h 2mi



Ψ∗ (r, t) ∇ Ψ(r, t) − Ψ(r, t) ∇ Ψ∗ (r, t)

 (748)

The expression for the probability current density, j(r, t), corresponds to the real function given by the velocity operator sandwiched between the wave functions   p ˆ Ψ(r) (749) j(r, t) = Real Ψ∗ (r) m The current density is real, and the real part is found by adding the expression and its complex conjugate, and then dividing by two. The proof that j(r, t) and ρ(r, t) satisfy the continuity equation starts from the Schr¨ odinger equation   ¯h2 ∂ 2 Ψ(r, t) = − ∇ + q φ(r, t) Ψ(r, t) (750) i¯ h ∂t 2m and its complex conjugate ∂ ∗ −ih ¯ Ψ (r, t) = ∂t



¯2 h − ∇2 + q φ(r, t) 2m



Ψ∗ (r, t)

(751)

Multiplying the first equation by Ψ∗ (r, t) and the complex conjugate equation by Ψ(r, t) and then subtracting it, one obtains   ¯h2 ∂ ∗ ∗ 2 2 ∗ Ψ (r, t) ∇ Ψ(r, t) − Ψ(r, t) ∇ Ψ (r, t) i¯ h Ψ (r, t) Ψ(r, t) = − 2m ∂t   ¯h2 = − ∇ . Ψ∗ (r, t) ∇ Ψ(r, t) − Ψ(r, t) ∇ Ψ∗ (r, t) 2m (752) which was to be proved. Thus, the total probability of finding the particle is unity, and remains unity for all times. This can be seen by integrating the continuity equation over a large fixed volume of space   Z Z ∂ 3 3 d r ∇ . j(r, t) = 0 (753) d r ρ(r, t) + ∂t 170

On using Green’s theorem to express the integral over the divergence of the current as an integral over the surface of the volume S, one has Z Z ∂ d3 r ρ(r, t) + d2 S . j(r, t) = 0 (754) ∂t If the current falls to zero at the boundary of the surface, one is left with Z ∂ d3 r ρ(r, t) = 0 (755) ∂t which, as the volume of integration is fixed, becomes Z d d3 r ρ(r, t) = 0 dt

(756)

Thus, one finds that the wave function that satisfies the Schr¨odinger equation remains normalized Z d3 r Ψ∗ (r, t) Ψ(r, t) = 1 (757) for all times. —————————————————————————————————— Example: The Probability Current Density of a Spherical Wave. Spherical Waves have wave functions of the form   exp ± i k r Ψ(r) = r

(758)

where k is a constant. This wave function is not normalizable and is independent of time. It should be interpreted as representing a steady state of a large number of beams of particles which are either diverging from or converging on the origin. The probability current density j(r) is calculated from   h ¯ j(r) = Ψ∗ (r) ∇ Ψ(r) − Ψ(r) ∇ Ψ∗ (r) 2mi

(759)

which involves the gradient of the spherical wave. The gradient is evaluated as     exp ± i k r 1 ∇Ψ(r) = eˆr ± ik − (760) r r which is purely radial. Hence, the current density is purely radial j = ±

¯ k 1 h eˆr m r2 171

(761)

and falls off as r−2 , with increasing r. The exponential phase factor in the wave functions cancel in the expression for the probability current density. The r−2 dependence of the probability current density is a consequence of continuity in the steady state, since it shows that the same number of particles, per unit time, pass through any spherical shell surrounding the origin, irrespective of the radius r of the shell. This is seen as follows: The number of particles passing through an infinitesimal area d2 S per unit time is given by the product j . d2 S = ±

¯ k eˆr h . d2 S m R2

(762)

where R is the radial distance of the surface element from the origin. On integrating this over the surface of a sphere of radius R, for which d2 S is also radially directed, one finds Z ¯h k dN − = d2 S . j = ± 4 π (763) dt m since, the surface area of the sphere is given by 4 π R2 . Therefore, the number of particles passing through the spherical surface is independent of R.

Spherical Waves

Figure 28: Outgoing Spherical Waves. The continuity condition is not satisfied at the origin and, depending on the sign of the probability current, the origin either acts as a source or sink where

172

particles are created or annihilated. ——————————————————————————————————

173

4 4.1

Applications of Quantum Mechanics Exact Solutions in One Dimension

One-dimensional problems are instructive, as they are often exactly soluble, and are relatively easy to solve. They also provide good illustrations of the principles of quantum mechanics. However, some phenomena found in one dimension may lead to conclusions that do not hold in higher dimensions. For example, an attractive potential in one dimension always leads to the formation of a bound state, but this does not remain true in higher dimensions. Nevertheless, we shall now examine some exactly soluble one-dimensional eigenstate problems.

4.1.1

Particle Confined in a Deep Potential Well

Let us consider the problem of a quantum mechanical particle, moving in one dimension. The particle is confined by a potential to move in a region of space where 0 < x < L, because the potential satisfies 0 V (x)

if L > x > 0

= ∞

otherwise

(764)

A classical particle with energy E < ∞ is forbidden to exist in the region of space outside the interval of length L. The particle’s motion √ consists of an orbit which moves back and forth with momentum p = ± 2 m E . The average position of the particle is given by x = L2 and the root mean squared deviation of the particle’s position is given by √L12 . The average value of the momentum is zero because it spends half the time moving forward with momentum + p and half the time moving backward with momentum − p. However, the root mean squared value of the particle’s momentum is exactly equal to p. Inside the interval L > x > 0, the energy eigenvalue equation takes the form h2 ∂ 2 ¯ − ΨE (x) = E ΨE (x) (765) 2 m ∂x2 which corresponds to the motion of a free particle. This equation is satisfied by the form     ΨE (x) = A exp + i k x + B exp − i k x (766) √ 2 m E . Within the region defined with arbitrary A and B, and p = h ¯k = by L > x > 0, the form of the solution appears to correspond to a linear superposition of momentum eigenstates, with momentum ± p.

174

Infinite Potential Well 120

V0

100

V(x)

80 60 40 20

V0 >> E > 0

0

E

-20 -1

-0.5

0

0.5

1

1.5

2

x/L

Figure 29: The potential of a deep potential well of depth V0 , where V0 → ∞. The zero of energy is chosen to be the bottom of the potential well. Outside the interval, the energy eigenvalue equation has the form   h2 ∂ 2 ¯ − + V ΨE (x) = E ΨE (x) 0 2 m ∂x2

(767)

where the limit V0 → ∞ should be taken. In this range only the form ΨE (x) = 0 satisfies this equation. This corresponds to the classical expectation that the particle should only be found within the interval of length L. The solution must be continuous for all values of x, even at the boundaries x = 0 and x = L. Thus, we require ΨE (x) to satisfy the boundary conditions ΨE (0) = ΨE (L) = 0

(768)

That is, the wave function inside the interval should match smoothly onto the solution outside the interval. This yields the two conditions A + B =0

(769)

at x = 0 and  A exp

 + ikL

 + B exp

175

 − ikL

=0

(770)

at x = L. The first condition yields A = − B and so inside the potential well the wave function is just      ΨE (x) = A exp + i k x − exp − i k x =

2 i A sin k x

(771)

for L > x > 0. The continuity condition at x = L then becomes ΨE (L)

= 2 i A sin k L = 0

(772)

Since we require A 6= 0, this places a condition on the allowed values of k which must satisfy k L = n π for integer values of n. The negative integer values of n give the same solution as the positive integers, and thus are not needed. The value n = 0 gives a trivial solution which doesn’t describe a particle at all. The solution n = 0 is not normalizable, and there is no probability for finding a particle in an interval ∆x even if L > x > 0. Thus, the allowed π where n is a positive integer greater than values of k are given by kn = n L zero. Therefore, the allowed values of the energy eigenvalue are also given by the discrete values En

= =

¯ 2 kn2 h 2m ¯h2 n2 π 2 2 m L2

(773)

The energy eigenvalues are discrete as the particle is trapped in a region of finite spatial extent. The corresponding energy eigenfunctions are given by Ψn (x) = C sin

nπx L

where C is an arbitrary complex number. The normalization condition Z +∞ dx | Ψn (x) |2 = 1

(774)

(775)

− ∞

reduces to

L

Z

dx | Ψn (x) |2 = 1

(776)

0

since Ψn (x) vanishes outside the region of integration. Using the form of Ψn (x) one finds that | C | is given by 1

= | C |2

Z

L

0

= | C |2

Z 0

L

nπx dx sin2 L   1 2nπx dx 1 − cos 2 L 176

1 = |C | 2 L = | C |2 2 2



L 2nπx x − sin 2nπ L

 x

= L

x = 0

(777)

Thus, the properly normalized wave functions are, " # r 2 nπx Ψn (x) = sin Θ(x) − Θ(x − L) L L

(778)

where Θ(x) is the Heaviside step function defined by Θ(x) Θ(x)

= 1 = 0

if x > 0 if x < 0

(779)

Eigenfunctions of the infinite potential well Ψ5(x) 0

Ψ4(x)

Ψn(x)

0

Ψ3(x)

0

Ψ2(x) 0

Ψ1(x)

0

0

0.2

0.4

0.6

0.8

1

x/L

Figure 30: The energy eigenfunctions Ψn (x) of a particle in a deep potential well. ˆ We These wave functions are the eigenfunctions of a Hermitean operator H. see that the eigenvalues are real En =

n2 π 2 ¯h2 2 m L2 177

(780)

and are non-degenerate. The eigenfunctions are orthogonal, as can be seen by directly evaluating the integral Z

+ ∞

dx Ψ∗m (x) Ψn (x)

2 L

=

− ∞

L

Z

dx sin 0

mπx nπx sin L L

  Z L 1 (m − n ) π x (m + n ) π x = dx cos − cos L 0 L L   1 sin( m − n ) π sin( m + n ) π = − π m − n m + n = δm − n (781) The overlap integral is zero for m 6= n and for m = n the wave functions are normalized to unity, as can be seen via invoking l’Hˆopital’s rule. Also one recognizes that functions in the finite interval L > x > 0 can, by the theory of finite Fourier series, be expanded in terms of the set of eigenfunctions. The choice of interval for the function which is to be expanded in a Fourier series, π x usually extends from − L to + L, and the series contains terms of cos n L in n π x addition to the terms sin L . However, if we insist that our wave functions Ψ(x) over the enlarged interval are odd, i.e. Ψ( − x ) = − Ψ( x ) then only the sin terms remain in the Fourier series expansion. That is, the eigenfunctions form a complete set, only in the interval L > x > 0 . The average values of the position of the particle in these energy eigenstates is given by Z

L

dx Ψ∗n (x) x Ψn (x)

x = 0

Z

=

2 L

Z

=

1 L

L

0

0

L

nπx dx x sin2 L   2nπx dx x 1 − cos L

However, the integral can be evaluated via Z Z ∂ dx x cos α x = dx sin α x ∂α   ∂ cos α x = − ∂α α   cos α x + x α sin α x = α2

(782)

(783)

Thus, we find that the average position is given by x =

178

L 2

(784)

Probability Density n=5 0

l Ψn(x) l

2

n=4 0

n=3 0

n=2 0

n=1 0 0

0.2

0.4

0.6

0.8

1

x/L

Figure 31: The quantum mechanical probability density P (x) = | Ψn (x) |2 for a particle in an energy eigenstate with energy eigenvalue E is compared with the classically calculated probability density for a particle with energy E. The classical probability density is calculated by averaging over the initial conditions for classical trajectories with energy E. In accordance with the correspondence principle, the probability densities should agree in the limit h ¯ → 0. which is exactly the same as the classical average. The mean squared deviation in the particle’s position is given by  2 Z L L ∆x2 = dx Ψ∗n (x) x − Ψn (x) 2 0   Z L 2 L2 nπx 2 = dx x − sin2 L 0 4 L    Z L 2 1 L 2nπx 2 = dx x − 1 − cos L 0 4 L    Z L 2 1 L 2nπx = dx x2 − 1 − cos L 0 4 L This integral is evaluated with the aid of the equality Z Z ∂2 2 dx x cos α x = − dx cos α x ∂α2 179

(785)

  ∂2 sin α x = − ∂α2 α   − 2 sin α x + 2 α x cos α x + α2 x2 sin α x = α3 (786) Thus, the mean squared deviation becomes ∆x2 =

L2 L2 − 12 2 n2 π 2

(787)

which for large enough n becomes identical to the classical value. This is another manifestation of the correspondence principle at work. These results correspond to a particle which has an average position half way along the allowed interval, and fluctuates back and forth with excursions proportional to the root mean squared deviation. The average value of the momentum is given by Z ∞ d p = −i¯ h dx Ψ∗n (x) Ψn (x) dx −∞ Z L 2 ¯h nπx d nπx = −i dx sin sin L L dx L 0   Z L 2 ¯h nπx d −i dx sin2 Θ(x) − Θ(x − L) L L dx 0 Z L 2πn¯ h nπx nπx = −i dx sin cos L2 L L 0   Z L 2 ¯h 2 n π x dx sin δ(x) − δ(x − L) −i L L 0 Z L πn¯ h 2nπx = −i dx sin L2 L 0   cos 2 n π − 1 πn¯ h = i L 2nπ = 0

(788)

The average value of the momentum of the quantum mechanical particle is zero, like the classical average of p. The root mean squared momentum is just given by Z ∞ d2 p2 = − ¯ h2 dx Ψ∗n (x) Ψn (x) dx2 −∞ Z L 2 ¯h2 n π x d2 nπx = − dx sin sin 2 L L dx L 0 180

− − = − − −

    Z L 2 ¯h2 nπx d nπx d 2 dx sin sin Θ(x) − Θ(x − L) L L dx L dx 0   Z L 2 ¯h2 n π x d2 dx sin2 Θ(x) − Θ(x − L) L L dx2 0 Z L n π x d2 nπx 2 ¯h2 dx sin sin 2 L L dx L 0    2 Z L 2 ¯h nπx d nπx 2 dx sin sin δ(x) − δ(x − L) L L dx L 0   Z L 2 ¯h2 2 n π x d dx sin δ(x) − δ(x − L) L L dx 0 Z L 2 π 2 n2 ¯ h2 nπx nπx dx sin sin L3 L L 0

=

+

=

π 2 n2 ¯h2 L2

(789)

In these calculations, the contributions from the discontinuities in the first and second derivatives of the Heaviside step functions vanish identically. Thus, the average momentum is zero just like for the classical particle, and the root mean n π squared deviation of the momentum is equal to p = h¯ L . This is similar to the classical particle, except that the r.m.s. momentum is quantized, that is, it only has discrete values. The momentum space wave function Φn (p) is given by r   Z L 1 2 nπx px sin exp − i Φn (p) = √ dx L L ¯ h 2π¯ h 0

(790)

which yields 

p L 2 h ¯

n π 2



   + n π sin i pL 2iL Φn (p) = √ +nπ (791)  2  2 exp − 2 ¯h π¯ hL p L nπ − h ¯ Hence, the momentum distribution function P (p) is given by P (p)

= | Φn (p) |2  2   n π nπ sin2 p2 L + h ¯ 2 4L = " 2  2 # 2 π ¯h p L nπ − h ¯

which is peaked at p = ±

n π h ¯ L .

181

(792)

Momentum distribution function 0.1

Pn(p)

0.08

0.06

n=4 0.04

0.02

0 -40

-30

-20

-10

0

10

20

30

40

pL/h

Figure 32: The momentum distribution function Pn (p) = | Φn (p) |2 of a particle in a deep potential well, with n = 4. The uncertainty in the momentum is given by ∆prms = uncertainty in the position is s  1 1 ∆xrms = L − 12 2 n2 π 2

h ¯ n π L

and the

(793)

The uncertainties satisfy the uncertainty principle, ∆xrms ∆prms ≥

4.1.2

¯ h 2

(794)

Time Dependence of a Particle in a Deep Potential Well

Consider a particle in an initial state given by an arbitrary wave function Ψ(x; 0), which satisfies the boundary conditions Ψ(0) = Ψ(L) = 0. Then on decomposing this into energy eigenstates via the expansion r X 2 nπx Ψ(x; 0) = Cn sin (795) L L n

182

one finds that the expansion coefficients are given by r Z L 2 nπx Cn = dx sin Ψ(x; 0) L 0 L

(796)

Then, the solution of the Schr¨odinger equation at time t is given by the expression  r X ¯h π 2 n2 2 nπx Ψ(x; t) = Cn exp − i t sin (797) 2 2mL L L n in which each term has a different time dependence.

4.1.3

Exercise 60

Find the time development of the initial wave function Ψ(x; 0) = δ( x − x0 ), of a particle in a box. Plot the time dependence of the probability density P (x; t) for finding the particle at position x at various times. Find the time dependence of the average value of the particle’s position x(t).

4.1.4

Particle Bound in a Shallow Potential Well

Let us consider the problem of a quantum mechanical particle moving in one dimension, in the presence of a potential given by − V0

if L > x > 0

V (x) = 0

otherwise

(798)

The energy of our particle is E < 0. Classically, the particle would confined by the potential to move in a region of space where 0 < x < L and the energy of the particle would be restricted by 0 > E > − V0 . Hence, the classical particle would be confined to the region near the origin. Inside the interval L form h2 ¯ − 2m which corresponds to the the form

> x > 0, the energy eigenvalue equation takes the ∂2 ΨE (x) = ( E + V0 ) ΨE (x) (799) ∂x2 motion of a free particle. This equation is satisfied by 

ΨE (x) = A exp

 + ikx

183

 + B exp

 − ikx

(800)

Shallow Potential Well 0.5

V(x) / V0

0

E

- V0 < E < 0 -0.5

- V0

-1

-1.5 -1

-0.5

0

0.5

1

1.5

2

x/L

Figure 33: A potential well V (x) of width L and depth − V0 . The bound state energy E is indicated by a horizontal line. p with arbitrary A and B, and p = h ¯ k = 2 m ( E + V0 ) . Within this region, the form of the solution appears to correspond to a linear superposition of momentum eigenstates, with momentum ± p. Outside the interval L > x > 0, the particle should satisfy the equation, −

¯h2 ∂ 2 ΨE (x) = E ΨE (x) 2 m ∂x2

(801)

which corresponds to the motion of a free particle but with negative energy. The kinetic energy instead of being positive must be negative, so this implies that we should consider imaginary values for the momentum. That is, we should try looking for solutions in which the momentum is given by p = i ¯h κ. In fact the energy eigenvalue equation is satisfied by the general form     ΨE (x) = Cα exp + κ x + Dα exp − κ x (802) p where h ¯ κ = 2 m | E | . The coefficients Cα and Dα are arbitrary and have to be separately determined in each region. The form of the solution corresponds to a linear superposition of momentum eigenstates, with imaginary value of the momentum ± i ¯h κ.

184

In the region where x > L the solution must be of the form of a decreasing exponential   ΨE (x) = D> exp − κ x for x > L (803) The term proportional to the coefficient C> has been discarded as the wave function must be square integrable, and since the term proportional to C> would diverge when x → ∞. The above wave function falls to zero at the boundary, x → ∞. In the region where x < 0, the solution must be of the form   ΨE (x) = C< exp + κ x for x < 0

(804)

as this tends to zero as x → − ∞. The term proportional to the coefficient D< has been discarded as the wave function must be square integrable. The above forms of the solution involve four (as yet) unknown constants A, B, C and D. These constant must be determined by the boundary conditions at x = 0 and x = L, and the normalization condition. As we shall presently see, the wave function and its derivative must be continuous at each of these points. Thus, there appears to be five conditions and only four unknowns. However, if we also regarded the energy E as a variable, then the above conditions may restrict the energy to discrete allowed values10 . The wave function should be continuous and also the first derivative should be continuous at both the points x = 0 and x = L. The need for the derivative of the wave function be continuous at these points can be seen by writing the potential as   V (x) = + V0

Θ( x − L ) − Θ( x )

(805)

and integrating the energy eigenvalue equation over a small interval of width 2  around either x = L or x = 0. Then L +  Z L+ Z L+ h2 ∂ ¯ + − ΨE (x) dx V (x0 ) ΨE (x0 ) = E dx ΨE (x0 ) 2 m ∂x L −  L −  L −  (806) Since the wave function is continuous, the term proportional to the energy has a magnitude of 2 E ΨE (L) . Also, as we are considering potentials V (x) which only have finite discontinuities and the wave function is continuous, the term 10 More precisely, the boundary conditions at | x | → ∞ are responsible for quantizing the energy. This can be shown by solving the eigenvalue equation for ΨE (x) (up to the normalization) for an arbitrary energy. If one enforces one boundary condition (say at x → − ∞) and integrates the equation, then the solution will diverge at the other boundary (x → + ∞) unless E is restricted to have certain discrete values.

185

Energy-dependence of trial energy-eigenfunctions of the attractive square well potential 2

E < E0

1.5

ΨE(x)

1

E = E0

0.5

E > E0

0 -1

-0.5

0

0.5

1

1.5

2

-0.5

x/L

Figure 34: The energy-dependence of trial energy eigenfunctions ΨE (x) for a shallow potential well. The solution is chosen to satisfy the boundary condition at x → − ∞ and is normalized to unity at x = 0. It is seen that the solution does not satisfy the boundary condition at x → ∞ unless E = E0 . This suggests that the discreteness of the bound state energy is caused by the boundary conditions at infinity. involving the integral of the potential has a magnitude of − V0  ΨE (L). In the limit  → 0, one finds ∂ ∂ ΨE (x) = ΨE (x) (807) ∂x ∂x x = L +  x = L −  Thus, the first derivative of the wave function must be continuous at the boundary. The continuity of the wave function at x = L yields the condition     ΨE (L) = A exp + i k L + B exp − i k L   = D exp − κ L

(808)

and the continuity of the first derivative yields      ∂ ΨE (L) = i k A exp + i k L − B exp − i k L ∂x 186

 = − κ D exp

 − κL

(809)

The continuity of the wave function at x = 0 results in the equation ΨE (0) = A + B = C and the continuity of the first derivative yields   ∂ ΨE (0) = i k A − B = +κC ∂x

(810)

(811)

Solving for A and B in terms of C, we find A = B

=

ik 2 ik 2

+ κ C ik − κ C ik

(812)

The two continuity conditions at x = L can be used to eliminate D and yields another equation, involving A and B      − κ A exp + i k L + B exp − i k L      = i k A exp + i k L − B exp − i k L (813) However, since A and B have already been determined in terms of C, and since C can be factored out (the magnitude of C is determined by the normalization condition), the above condition restricts the possible values of the energy E. Hence, we obtain the secular equation which determines the eigenvalues !   ( i k − κ )2 exp + 2 i k L = (814) ( i k + κ )2 On taking the square root, we can rewrite this as   ( i k − κ )2 exp i k L = ± ( k 2 + κ2 )

(815)

which can be rewritten as an equation for the ratio of the imaginary and real parts, in which case the ± sign drops out tan k L =

2kκ k 2 − κ2

187

(816)

which we shall refer to later on. What we really need to do is to take the fourth root of the equation (814), which yields the two equations   (ik − κ) exp i k L / 2 = ± i p (817) ( k 2 + κ2 ) and

 exp

 ikL/2

(ik − κ) = ± p ( k 2 + κ2 )

(818)

which correspond to the four fourth roots of unity. On taking the ratio of the imaginary to the real parts, these two equations can be re-written as two conditions kL 2 kL k cot 2

k tan

= +κ = −κ

(819)

The above two conditions determine the bound state energies E as the wave vector k and the value κ are related to the energy through 2mE ¯h2 2 m ( E + V0 ) ¯h2

κ2

= −

k2

=

(820)

The conditions on k can be solved graphically for the bound states, by squaring the equation. For the first condition, as both k and κ are positive, squaring introduces spurious solutions, corresponding to regions where k tan k 2L is negative. We then plot 2 m V0 κ2 = − k2 (821) ¯h2 as a function of k, on the same graph as the function k 2 tan2

kL 2

For positive V0 we have a parabola that intersects the k axis at k =

(822) q

2 m V0 . h ¯2 2 n π for L

The function k 2 tan2 k 2L is never negative, but falls to zero at k = 1 ) π integer n and diverges to infinity at k = ( 2 n + . Thus, for k in the range L q ( 2 n + 1 ) π 2 m V0 0 < k < , one has roots near k = . Since tan k 2L L h ¯2

1 ) π is negative in each alternate segment 2 nL π > k > ( 2 n − where the L slope of our function is negative, the solution in these regions are spurious. The solutions of our condition corresponds to the intersections of our function with

188

Self Consistency Equation 200

k2tan2kL/2 150

k2 cot2kL/2

100

2mV0/h2- k2

50

0 0

1

2

3

4

5

6

7

8

kL

Figure 35: Graphical solution of the equations determining the bound state energies of the shallow potential well. The k values where the inverted parabola 2 kL 2 intersects the curves k 2 tan2 kL 2 or k cot 2 represents possible allowed wave vectors k. As explained in the text, there is a maximum of one allowed solution π for k in each interval of L , as the second solution is spurious. the parabola, but only in the intervals function has a positive slope.

( 2 n + 1 ) π L

> k >

2 n π L

where our

The energy eigenstates satisfying the second condition in eqn(819) can be found by a similar procedure. However, we should note that in this case k is real and − κ is negative. Thus, the roots of the second condition should lie in the intervals where tan k 2L is negative, and the intersections in the regions ( 2 n + 1 ) π > k > 2 nL π are spurious and should be discarded. Thus, the L regions of k which contain bound states are interchanged for the two conditions. Let us note that for large enough V0 the bound states correspond to the set of k values given by nπ k ≈ (823) L if the values of k are sufficiently small and are limited by ¯ 2 k2 h  V0 2m

189

(824)

In this case, the eigenvalues are related to the eigenvalues of the infinite potential well. Since the values of k are found to be discrete, the energy eigenvalues are also discrete. The allowed values of E are found to be in the range 0 ≥ E ≥ − V0 . The only remaining undetermined constant is the value of C which can be found by normalizing the wave function.

Bound state wave functions for a shallow potential well 1.5

Ψ2(x)

Ψ0(x)

1

Ψn(x)

0.5 0 -1

-0.5

0

0.5

1

1.5

2

-0.5

Ψ1(x)

-1 -1.5

x/L

Figure 36: Bound state wave functions Ψn (x) for a shallow potential well. ——————————————————————————————————

4.1.5

Exercise 61

A particle of mass m is confined to move in a one-dimensional potential V (x) where V (x) → ∞ V (x) = − V0 V (x) = 0

x < 0 0 < x < a a < x

Derive the equation for the bound state energy. Find the minimum value of V0 which will produce a bound state.

190

(825)

An asymmetric potential well 0.5

V(x) / V0

0

E

- V0 < E < 0

-0.5

- V0

-1

-1.5 -1

-0.5

0

0.5

1

1.5

2

x/L

Figure 37: A one-dimensional potential V (x), which excludes the particle from the region x < 0. The minimum value of the potential is − V0 . The potential V0 has to exceed a minimum value if the particle is to be bound. ——————————————————————————————————

4.1.6

Solution 61

The bound state wave function is given by Ψ(x)

= 0

for x < 0

where V (x) → ∞. In the region of positive x, the energy eigenvalue equation is given by h2 ∂ 2 Ψ ¯ + V0 Ψ(x) = E Ψ(x) (826) − 2 m ∂x2 inside the potential well where a > x > 0, and −

¯ 2 ∂2Ψ h = E Ψ(x) 2 m ∂x2

outside the potential well where x > a. This has the solution     Ψ(x) = A exp + i k x + B exp − i k x

191

(827)

(828)

Bound state wave function with minimal energy

Ψ0(x)

0.15

Ψ0(x)

0.1

0.05

0 -1

0

1

2

x/L

Figure 38: The form of the bound state wave function, for a value of the potential V0 which is just sufficently strong enough to create one bound state. where the energy eigenvalue is given by E = − V0 +

¯2 2 h k 2m

(829)

Since the wave function has to be continuous at x = 0, one has A + B = 0

(830)

Thus, the solution inside the well is of the form Ψ(x) = 2 i A sin k x

(831)

Outside the potential well, the eigenvalue equation has the solution of the form     Ψ(x) = C exp − κ x + D exp + κ x (832) However, as the wave function must vanish as x → ∞ then one has D = 0. Thus, the wave function for x > a is simply given by   Ψ(x) = C exp − κ x (833)

192

and the energy eigenvalue is given by E = −

¯2 2 h κ 2m

(834)

Continuity at x = a yields  2 i A sin k a = C exp

 − κa

(835)

and continuity of the derivative yields  2 i k A cos k a = − κ C exp

 − κa

(836)

Hence on dividing the above equations, one has k cot k a = − κ

(837)

On squaring this equation, one has 2 m V0 − k2 ¯h2

k 2 cot2 k a = or k2 =

2 m V0 sin2 k a ¯h2

(838)

(839)

Thus, 2 m ( V0 + E ) 2 m V0 = sin2 k a h2 ¯ ¯2 h

(840)

If a bound state is just formed, it has the maximum bound state energy which is zero, E = 0. Hence, in this case, the above equation simplifies to 1 = sin2 k a

(841)

Thus, the minimum value of k is found as π 2a

k =

(842)

and since E = 0, one has 0 = − V0 +

¯ 2 π2 h 2 m 4 a2

(843)

Hence, the minimum value of V0 needed to produce a bound state is given by V0 =

¯ 2 π2 h 8 m a2

193

(844)

Alternatively, the result can be derived from the condition that the kinetic energy inside the well and outside the well must be minimized. The wave function inside the well is oscillatory, but must have the minimum number of oscillations. The wave function outside the well must decay slowly, therefore, the decay constant for the wave function for x > a is given by κ = 0. Thus, the matching condition for the derivative of the wave function shows that the slope must be zero at the edge of the potential well x = a and, since E = 0. Since the wave function has a maximum at x = a, one has k cos k a = 0 from which one has k =

π 2 a

E

(845)

. Hence, as = −

¯h2 2 κ 2m

¯2 2 h k (846) 2m the minimum value of the potential needed to produce a bound state is given by ¯h2 π 2 V0 = (847) 8 m a2 = − V0 +

——————————————————————————————————

4.1.7

Scattering from a Shallow Potential Well

The shallow potential well is described by V (x) = − V0 for 0 < x < L V (x) = 0 otherwise

(848)

Classically, the energy E = 0 separates the bound states, which have E < 0, and the scattering states, which have E > 0. We shall now consider states where the energy E is greater than zero so that the particle may travel to infinity. The allowable forms for the wave function in the three regions are     Ψ(x) = A exp + i k x + B exp − i k x 0 > x     Ψ(x) = C exp + i k 0 x + D exp − i k 0 x L > x >0     Ψ(x) = F exp + i k x + G exp − i k x x > L (849) 194

Scattering from a Potential Well 0.5

E>0

V(x) / V0

0

-0.5

-1

- V0

-1.5 -1

-0.5

0

0.5

1

1.5

2

x/L

Figure 39: Scattering from a shallow potential well V (x) of depth − V0 . The energy of the scattering states E is > 0, so classically the particle is free to move to infinity. 2

2

2

02

k k and E + V0 = h¯2 m where the energy eigenvalue is given by E = h¯2 m However, as we only have four matching conditions, the six coefficients are not determined, and k is a continuous real variable. We can choose one coefficient to set the normalization of the wave function. The other coefficient can be chosen as desired. We shall choose the coefficient G to be zero. This represents a beam of particles being produced at x → − ∞ with momentum p falling incident on the potential at x = 0. Some fraction of the beam is reflected back towards x → − ∞, with momentum − p and a certain fraction being transmitted to x → + ∞ with momentum p. The vanishing of G corresponds to the absence of a source of particles, with momentum − p, at x → ∞.

The matching conditions at x = 0 yield A + B = C + D k ( A − B ) = k0 ( C − D )

(850)

The matching conditions at x = L are given by       0 0 C exp + i k L + D exp − i k L = F exp + i k L

195

k0



 C exp

+ i k0 L



 − D exp

− i k0 L



 = k F exp

 + ikL (851)

This pair of equations can be solved for C and D in terms of the amplitude of the transmitted beam F ,    0    k + k 0 F exp i k L C exp + i k L = 2 k0    0    k − k 0 D exp − i k L = F exp i k L (852) 2 k0 whereas the amplitudes of the by  k A =  k B =

incident beam A and reflected beam B are given    k − k0 + k0 C + D 2k 2k    − k0 k + k0 C + D 2k 2k

(853)

Thus, we find     2 k k 0 cos k 0 L − i ( k 2 + k 02 ) sin k 0 L A exp − i k L = F 2 k k0     ( k 02 − k 2 ) i sin k 0 L B exp − i k L = F (854) 2 k k0 The amplitude of the incident beam is assumed to be known, as it can be controlled by the particle accelerator. Thus, we shall express the amplitudes of the transmitted and reflected beams as ratios with respect to the incident beam. The ratio of the transmitted beam to the incident beam is given by     2 k k 0 exp − i k L F = (855) A 2 k k 0 cos k 0 L − i ( k 2 + k 02 ) sin k 0 L

The ratio of the amplitudes of the reflected beam to the incident beam is given by   B ( k 02 − k 2 ) i sin k 0 L = (856) A 2 k k 0 cos k 0 L − i ( k 2 + k 02 ) sin k 0 L The amplitude of the reflected beam is proportional to k 02 − k 2 and thus, is proportional to V0 , and the denominator varies as E for large E. The denominator is non-zero when both k and k 0 are real. This is because the magnitude 196

of the denominator is given by the sum of two squares 4 k 2 k 02 cos2 k 0 L + ( k 2 + k 02 )2 sin2 k 0 L

(857)

and when the sine factor is zero, the cosine factor is positive. Thus, there are no simultaneous zeros. In the above expressions k and k 0 are continuous variables and so the energy eigenvalues have a continuous spectrum in the range E ≥ 0. The intensities of the reflected and transmitted beams add to equal the intensity of the incident beam, | A |2 = | B |2 + | F |2

(858)

This is a necessary condition for the probability to be conserved. As the probability does not build up with time at any point in space, but is in a steady state, the probability currents entering any region of space must be equal to those leaving. As the magnitude of the velocities outside the potential well are simply given by h¯mk , the equality of incoming flux with the outgoing flux reduces to the above statement about the intensities of the beams. The reflection coefficient is defined as the ratio R =

| B |2 | A |2

(859)

which takes into account that the magnitude of the velocity of the incident beam is always the same as that of the reflected beam, as they are moving in the same region of space, and thus experience the same potential. The reflection coefficient is given by   ( k 02 − k 2 )2 sin2 k 0 L R(k) = (860) 4 k 2 k 02 cos2 k 0 L + ( k 2 + k 02 )2 sin2 k 0 L We see that the reflection coefficient has a numerator proportional to V02 , and the denominator grows like E 2 for sufficiently large E. It should be noted that, in this case, quantum mechanics produces results which are completely different from classical mechanics as, in one dimension, a classical particle is not reflected by an attractive potential well. The transmission coefficient is defined as the ratio T =

kt | F |2 ki | A |2

(861)

in which kt is the magnitude of the wave vector of the transmitted beam F exp[ i kt x ] and ki is the magnitude of the wave vector for the incident 197

Reflection Coefficient for a beam of particles incident on a shallow potential well 1

R(k)

0.8 0.6 0.4 0.2 0 0

2

4

6

8

10

12

kL

Figure 40: The reflection coefficient R(k) for a beam of particles of momentum ¯ k incident on a shallow potential well of depth − V0 and width L. Note that, h due to the Ramsauer effect, the reflection coefficient falls to zero periodically. beam A exp[ i ki x ]. In our examples ki and kt have the same value and, therefore, cancel. In the more general case, the velocities have to be included, and then the reflection and transmission coefficients will satisfy the equation, R + T = 1

In the above example, the transmission coefficient is evaluated as   4 k 2 k 02 T (k) = 4 k 2 k 02 cos2 k 0 L + ( k 2 + k 02 )2 sin2 k 0 L

(862)

(863)

It is seen that the beam is totally transmitted whenever sin k 0 L = 0. The transmission coefficient shows oscillations with increasing k 0 . The minimum in the reflection coefficient and the corresponding maximum in the transmission coefficient is known as the Ramsauer effect11 . It is seen in the scattering of low-energy electrons from inert gas atoms. —————————————————————————————————— 11 C.

Ramsauer, Ann. Phys. (Leipzig) 64, 546 (1921).

198

4.1.8

Exercise 62

Consider a particle of mass m moving in one dimension in a potential given by 0 V (x)

if

x < 0

= − V0

if

x > 0

(864)

Consider the energy eigenstate which corresponds to a beam of particles of momentum − p incident on the potential step from x → ∞, find the amplitude of the reflected and transmitted waves, for E > 0. Also find the reflection and transmission coefficients and show that they add up to unity. ——————————————————————————————————

4.1.9

Solution 62

The incident beam is in the region x > 0 and is represented by the wave travelling along the negative x direction   A exp − i k 0 x (865) In addition, for positive x, one expects a reflected wave of amplitude B that travels along the positive x direction. The form of the solution for x > 0 is given by     0 0 Ψ(x) = A exp − i k x + B exp + i k x (866) On substituting the trial form of the solution in the energy eigenvalue equation for x > 0 h2 ∂ 2 ¯ − Ψ(x) − V0 Ψ(x) = E Ψ(x) (867) 2 m ∂x2 where we have used V (x) = − V0 for x > 0, one obtains a relation between the energy eigenvalue and k 0 , ¯ 2 02 h k − V0 = E 2m

The solution for x < 0 corresponds to the transmitted beam   Ψ(x) = C exp − i k x

199

(868)

(869)

which on substituting into the energy eigenvalue equation −

¯ 2 ∂2 h Ψ(x) = E Ψ(x) 2 m ∂x2

leads to

¯2 2 h k = E 2m

(870)

(871)

The matching conditions at x = 0 yield C kC

= A + B = k0 ( A − B )

(872)

These can be solved to yield the amplitude of the reflected and transmitted waves  0  B k − k = A k0 + k   C 2 k0 = (873) A k0 + k as ratios to the amplitude of the incident wave. The transmission coefficient, T , is given by 2 k C T = k0 A 4 k k0 = (874) ( k + k 0 )2 where the factor k / k 0 represents the change in the flux of particles due to the difference in the velocities of the transmitted incident beam. The reflection coefficient R is given by B 2 R = A ( k 0 − k )2 = (875) ( k + k 0 )2 Hence, we find that reflection and transmission coefficients add R + T

( ( ( = ( = 1 =

k0 k k k

− k + k0 + k0 + k0

)2 4 k k0 + 2 ) ( k + k 0 )2 )2 )2 (876)

200

to yield unity. ——————————————————————————————————

4.1.10

Exercise 63

By considering the solution of the Schr¨odinger equation with the initial state described by the energy eigenstate found in the previous example, show that the system is in a steady state. Find the probability current density in the two regions, and show that it satisfies the continuity condition, when E > 0. Also, show that the continuity condition is satisfied in the two regions when E < 0. ——————————————————————————————————

4.1.11

Solution 63

The probability density ρ(x, t) for finding a particle in the region x > 0 is given by 2 ρ(x, t) = Ψ(x, t)   2     2 i 0 0 = A exp − i k x + B exp + i k x exp − E t ¯ h     2 2 ∗ 0 ∗ 0 = | A | + | B | + A B exp − 2 i k x + A B exp + 2 i k x "  #   0  2 02 k + k k − k = | A |2 2 + 2 cos 2 k 0 x (877) ( k + k 0 )2 k0 + k which is time independent, but exhibits oscillations due to the interference between the incident and reflected wave. The probability density ρ(x, t) for finding a particle in the region x < 0 is given by 2 ρ(x, t) = Ψ(x, t)     2 i = C exp − i k 0 x exp − E t ¯ h = | C |2 = | A |2

4 k 02 ( k + k 0 )2

which is spatially uniform and time independent.

201

(878)

The probability current density jx (x, t) is given by # "   i ¯h ∂ ∗ Ψ(x, t) jx (x, t) = Real Ψ (x, t) − m ∂x

(879)

For x > 0, one finds the " # ¯h k 0 2 2 jx (x, t) = Real − (|A| − |B| ) m "     # ¯h k 0 + Real − ( A B ∗ exp − 2 i k 0 x − B A∗ exp + 2 i k 0 x ) m " # h k0 ¯ 2 2 = Real − (|A| − |B| ) m (880) as the last term is purely imaginary. This represents the net flux along the positive direction, which is the difference between the transmitted flux and the incident flux. Note that the ratio of these two fluxes is (apart from a sign) just the reflection coefficient R. The total probability current density for x > 0 is evaluated as   ¯h 0 4 k k0 jx (x, t) = − k | A |2 (881) m ( k + k 0 )2

For x < 0 one finds jx (x, t) = −

¯ k h | C |2 m

(882)

On evaluating the probability current density for x < 0, one finds that the current density is time independent and spatially uniform. The current density has a value of   ¯h 4 k 02 jx (x, t) = − k | A |2 (883) m ( k + k 0 )2 Note that this just represents the flux corresponding to the transmitted beam. The ratio of the transmitted flux to the incident flux is just the transmission coefficient T . Since, the probability current density has the same value for x > 0 and x < 0 it is independent of x. The probability density is also independent of t. Thus, the continuity condition ∂ρ(x, t) ∂jx (x, t) + = 0 ∂t ∂x 202

(884)

is trivially satisfied. ——————————————————————————————————

4.1.12

The Threshold Energy for a Bound State

We have examined the shallow potential well and found energy eigenstates in the energy regions, E > 0, which corresponded to scattering states with continuous energy, and the energy region 0 > E > − V0 which had bound states with discrete energy eigenvalues. This raises the question as to whether there are any states with energy E less than − V0 . Classically, it is impossible to have states with an energy lower than the minimum of the potential energy as the only other contribution to the total energy, apart from the potential, is the kinetic energy which is positive definite. We shall now prove that it is impossible to have quantum states with energy eigenvalues that are lower than the minimum value of the potential. If a bound state exists, then | Ψ(x) |2 must decay exponentially as | x | → ∞ in order that the wave function be normalized. This means that | Ψ(x) |2 must have at least one maximum. This means that ∂ | Ψ(x) |2 = 0 ∂x

(885)

and

∂2 | Ψ(x) |2 < 0 ∂x2 at the value of x = x0 where the maximum is located. We shall start with the energy eigenvalue equation   h2 ∂ 2 Ψ ¯ − + V (x) − E Ψ = 0 2 m ∂x2

(886)

(887)

and pre-multiply by the complex conjugate wave function Ψ∗ and add this to the complex conjugate of the energy eigenvalue equation post-multiplied by Ψ. The result is     h2 ¯ ∂2Ψ ∂ 2 Ψ∗ − Ψ∗ + Ψ + 2 V (x) − E | Ψ |2 = 0 (888) 2m ∂x2 ∂x2 We can re-write the above equation as  2    h2 ¯ ∂ ¯h2 2 2 − | Ψ | + 2 V (x) − E | Ψ | = − 2m ∂x2 m

∂Ψ 2 ∂x (889)

This must hold at all values of x, including x0 . But at the maximum we have ∂ | Ψ(x0 ) | = 0 ∂x 203

(890)

and

∂2 | Ψ(x0 ) |2 < 0 ∂x2 Hence, we find that the equality  2    h2 ¯ ∂ 2 − | Ψ(x ) | = 2 E − V (x ) | Ψ(x0 ) |2 0 0 2m ∂x2

(891)

(892)

holds at x0 . Since the first term is positive, the second term must also be positive. Furthermore, as | Ψ(x0 ) |2 ≥ 0, we must have E ≥ V (x0 )

(893)

By definition, V (x0 ) is greater than or equal to the minimum value of V (x). Therefore, the bound state energy E must be greater than or equal to the minimum value of the potential.

4.1.13

Transmission through a Potential Barrier

Let us now consider a particle with energy E > 0 moving in a potential which now has the form V (x) V (x)

= + V0 if L > x > 0 = 0 otherwise

(894)

We shall consider states in the energy range where V0 > E > 0. In this energy range, a classical particle would be forbidden to exist in the region L > x > 0 but could be found in the other regions of space. A specific form of the solution of the Schr¨odinger equation can be written as     Ψ(x) = A exp + i k x + B exp − i k x x < 0     Ψ(x) = C exp + κ x + D exp − κ x L > x > 0   Ψ(x) = F exp + i k x x > L (895) where k is related to the energy eigenvalue E via 2mE ¯h2 2 m ( V0 − E ) ¯h2

k2

= +

κ2

=

(896)

This form is appropriate to the case where a beam of particles of momentum p = h ¯ k is incident from x → − ∞ onto the potential barrier. A portion of 204

Potential Barrier 1.5

V0 V(x) / V0

1

V0 > E > 0

E

0.5

0

-0.5 -1

-0.5

0

0.5

1

1.5

2

x/L

Figure 41: A potential V (x) with a barrier of height V0 and width L. the beam will be reflected back to x → − ∞ and a portion will be transmitted to x → ∞. The continuity condition at x = 0 can be solved to yield     +κ + ik −κ + ik A = C + D 2ik 2ik     −κ + ik +κ + ik B = C + D (897) 2ik 2ik The continuity condition to  κ C =  κ D =

at x = L yields C and D in terms of F . This leads      + ik F exp i k L exp − κ L 2κ      − ik F exp i k L exp + κ L 2κ

(898)

On substituting the expressions for C and D into the previous expressions for A and B, one obtains    exp i k L  A = 4 i k κ cosh κ L + 2 ( k 2 − κ2 ) sinh κ L F 4ikκ

205

 exp B

=

 ikL

 2(k

4ikκ

2



2

+ κ ) sinh κ L

F

(899)

The ratio of the amplitude of the transmitted wave to the incident wave is F / A which is found as   4 i k κ exp − i k L F  =  (900) A 2 2 4 i k κ cosh κ L + 2 ( k − κ ) sinh κ L whereas the ratio of the amplitude of the reflected wave to the incident wave is B / A where   2 ( k 2 + κ2 ) sinh κ L B  =  (901) A 4 i k κ cosh κ L + 2 ( k 2 − κ2 ) sinh κ L These expressions are obviously related to the amplitudes for reflection from and transmission through a shallow potential well, via the analytic continuation k0 → i κ

(902)

The transmission coefficient T (k) is evaluated as T (k)

F 2 = A =

4 k 2 κ2  4

k2

κ2

2

cosh κ L + (

k2



κ2

)2

2



(903)

sinh κ L

  For large κ L, we have cosh κ L ' 12 exp + κ L and sinh κ L '  1 2 exp + κ L . Thus, we find that transmission coefficient decreases exponentially with increasing L. The amplitude of the transmitted wave is exponentially suppressed, and the leading exponential term is     F 4ikκ ≈ exp − κ L exp − i k L (904) A ( k + i κ )2 The phenomenon in which particles are transmitted through a potential barrier (known as quantum tunnelling) is due to the exponentially small probability of finding the particles in the region of space where they are classically forbidden to exist due to energetic reasons. In the classically forbidden regions, the classical 206

Transmission through a potential barrier

Real [ Φk(x) ]

R

-2

-1

T

0

1

2

3

4

x/a

Figure 42: Transmission Through a Potential Barrier. The real part of the wave function Φk (x) is plotted as a function of x. A beam of particles is emitted from a source located at x → − ∞ and falls incident on the potential barrier. The barrier extends from x = 0 up to x = a, and its extremities are marked by the red vertical lines. The height of the barrier V0 is greater than the kinetic energy E of the incident particles, and hence is impenetrable to a beam of classical particles of energy E. Most of the incident beam is reflected from the potential barrier. The remaining part of the beam tunnels through the potential barrier and is transmitted to infinity. Therefore, some of the particles quantum mechanically tunnel through the classically forbidden region. kinetic energy would have to be negative. In the limit of large κ L, within the approximation of leading exponential terms, the reflected wave has a relative amplitude of unity ( k 2 + κ2 ) B ≈ (905) A ( k + i κ )2 as is needed if the probability current is to satisfy the continuity equation. ——————————————————————————————————

4.1.14

Exercise 64

Construct the probability current density in the three regions and show that the continuity equations is satisfied in each region, including the region where

207

Transmission Coefficient for a Potential Barrier 1.2 1

T(k)

0.8 0.6 0.4 0.2 0 0

2

4

6

8

10

12

kL

Figure 43: The transmission coefficient T (k) of a beam of particles, of wave vector k, incident on a potential barrier of height V0 . For energies in the range V0 > E > 0, where a beam of classical particles would be totally reflected, the quantum mechanical particles have an exponentially small probability of being transmitted. tunnelling occurs. ——————————————————————————————————

4.1.15

Solution 64

Since the states are energy eigenstates, the probability density is time independent. For x < 0, the probability density      ρ(x, t) = | A |2 + | B |2 + A B ∗ exp i 2 k x + B A∗ exp − i 2 k x (906) is oscillatory. For L > x > 0, the probability density is also non-uniform       2 2 ∗ ∗ ρ(x, t) = | C | exp + 2 κ x + | D | exp − 2 κ x + C D + DC (907) 208

For x > L, the probability density is spatially uniform ρ(x, t) = | F |2

(908)

The probability current density jx (x, t) in the region x < 0 is given by   ¯h k 2 2 jx (x, t) = |A| − |B| (909) m In the barrier region, the probability current density is given by   ¯h κ jx (x, t) = − i C D∗ − C ∗ D m

(910)

for L > x > 0. This quantity is only finite if the coefficients C and D are complex. Since the coefficients are determined via the matching conditions, the probability current density is only finite if the wave function shows periodic oscillations in some regions of space. In the last region, the probability current density is also spatially uniform jx (x, t)

=

¯ k h | F |2 m

(911)

for x > L. On evaluating the probability current density, it is found to be given by jx (x, t) =

4 k 2 κ2 ¯h k | A |2 (912) m 4 k 2 κ2 cosh2 κ L + ( k 2 − κ2 )2 sinh2 κ L

and is uniform throughout space. Since the probability current density is constant and the probability density is time independent, the continuity equation ∂ρ ∂jx + = 0 ∂t ∂x

(913)

is trivially satisfied. ——————————————————————————————————

4.1.16

The Double Well Potential

Consider a particle of mass m moving in one dimension in the presence of a double well potential V (x) = ∞ V (x) = 0 V (x) = V0

for | x | > ( a + b ) for ( a + b ) > | x | > ( b − a ) otherwise 209

(914)

The double well potential

V(x)

V0

En

En -(b+a)

-(b-a)

x

(b-a)

(b+a)

Figure 44: The double well potential. The potential V (x) is a symmetric function of x. The parity operator Pˆ is defined as the operator which, when acting on an arbitrary function ψ(x), has the effect Pˆ ψ(x) = ψ(−x) (915) On applying the parity operator twice, it is seen that it satisfies the relation Pˆ 2 = Iˆ

(916)

The eigenfunctions φp (x) of the parity operator satisfy the equations Pˆ φp (x) = p φp (x)

(917)

where p is the parity eigenvalue. Because of the relation expressed by eqn(916), the eigenvalues of the parity operator are restricted to have values ± 1. Since the potential is an even function of x, the Hamiltonian commutes with the parity operator ˆ ] = 0 [ Pˆ , H (918) so that the parity is a conserved quantity. In other words, parity is a constant of motion. Since the Hamiltonian commutes with the parity operator, one can find ˆ and Pˆ . We shall seek such solutions for which simultaneous eigenstates of H E < V0 . In this case, classical particles are excluded from traversing the region containing the potential barrier.

210

The even parity eigenfunctions can be expressed in the form φ(x) = A sin k ( x − a − b ) φ(x) = B cosh κ x

for ( a + b ) > x > ( b − a ) for ( b − a ) > | x | (919)

where the energy eigenvalue is related to k and κ via E =

¯ 2 κ2 h ¯ 2 k2 h = V0 − 2m 2m

(920)

Since φ(x) is an even parity eigenfunction φ(−x) = φ(x), the form of the solution for − ( b − a ) > x > − ( b + a ) is given by φ(x)

= − A sin k ( x + a + b )

for − ( b − a ) > x > − ( b + a ) (921)

The above form of the solution φ(x) identically satisfies the boundary conditions at the hard walls x = ± ( b + a ). The remaining matching conditions at x = ± ( b − a ) lead to the self-consistency equation k cot 2 k a = − κ tanh κ ( b − a )

(922)

This equation together with the relation k 2 + κ2 =

2 m V0 ¯h2

(923)

can be solved graphically. The solutions determine the quantized values of k. The odd parity eigenfunctions can be expressed as φ(x) φ(x)

= A sin k ( x − a − b ) = B sinh κ x

for ( a + b ) > x > ( b − a ) for ( b − a ) > | x | (924)

The form of the solution for negative values of x can be obtained from those of positive values of x by noting that they have odd parity φ(−x) = − φ(x). If the above forms of φ(x) are to satisfy the differential equation, the energy eigenvalue must be related to k and κ via eqn(920). The above form of the solution φ(x) automatically satisfies the boundary conditions at the hard walls x = ± ( b + a ). The matching conditions for the odd parity eigenfunctions at x = ± ( b − a ) lead to the equation k cot 2 k a = − κ coth κ ( b − a )

(925)

The quantized values of k are given by the solutions of the above self-consistency equation.

211

Consistency conditions 5

k cot 2ka 0

−κ tanh κ(b-a) -5

−κ coth κ(b-a)

-10 0

0.5

1

2ka /π

1.5

2 2

2 2

2(2mV0a /h π ) 1/2

Figure 45: The graphical solution of the self consistency conditions for the double potential well. The right hand side for odd parity solutions is shown by the dashed line, the right hand side for even parity solutions is shown by the solid line. Graphical solution of the even and odd parity self consistency equations shows that, for large barrier heights or large barrier widths such that tanh κ( b − a ) ∼ 1

(926)

the eigenvalues occur in pairs which are almost degenerate. The pairs of energies have magnitudes which are close to the bound state energies of the isolated wells, so k cot 2 k a ∼ − κ (927) The splitting ∆E between the lowest pairs of eigenvalues is exponentially small, and is approximated by   h2 π 2 2 ¯ ∆E ∼ exp − 2 κ ( b − a ) (928) 2 m a2 κ a The tunnel splitting between the pairs of energy levels increases for the higher energy levels.

212

Energy eigenfunctions of the double well potential

Ψn(x)

Ψ0(x)

-(b+a)

-(b-a)

(b+a)

(b-a)

Ψ1(x)

x

Figure 46: The lowest pair of energy eigenfunctions for the double well potential. 4.1.17

The delta function Potential

Let us consider a beam of particles of mass m and momentum ¯h k scattering from an attractive potential V (x) localized at the origin V (x) = − V0 a δ(x)

(929)

where V0 a is a measure of the strength of the potential. In the strength, V0 has units of potential and a has units of length. The energy eigenvalue equation becomes   h2 ∂ 2 ¯ − − V0 a δ(x) φk (x) = E φk (x) 2 m ∂x2 We shall look for a solution which is of the form     φk (x) = A exp + i k x + B exp − i k x

(930)

(931)

in the region 0 > x and has the form  φk (x) = F exp

213

 + ikx

(932)

Delta function potential 2 0

E0

V(x) = - V0 a δ(x)

V(x)

-2 -4 -6 -8 -10 -12 -14 -2

-1

0

x/a

1

2

Figure 47: The attractive delta function potential. The bound state energy is shown by a horizontal dashed line. for x > 0. These forms satisfy the energy eigenvalue equation in these two regions if ¯h2 k 2 E = (933) 2m since, the potential vanishes identically in these two separate regions. Again, the form is appropriate for discussion of the scattering of a beam of particles with energy E > 0 and momentum ¯h k incident on the potential from x → − ∞. The matching conditions at x = 0 are affected by the presence of the non-zero potential. The continuity of the wave function yields A + B = F

(934)

The discontinuity in the first derivative is produced by the infinite potential at the origin. Let us integrate the energy eigenvalue equation from x = −  to x = +  and let  → 0. Then, as E and φk (0) are finite, in the limit  → 0 one has   Z + Z + h2 ∂ 2 ¯ dx − − V0 a δ(x) φk (x) = E dx φk (x) 2 m ∂x2 −  −  +  h2 ∂ ¯ − φk (x) − V0 a φk (0) = E 2  φk (0) (935) 2 m ∂x − 

214

which reduces to −

+ ¯2 ∂ h φk (x) 2 m ∂x −



= V0 a φk (0)

(936)



This yields the second condition     h2 ¯ ¯h2 − ikF + ikA − ikB = V0 a F 2m 2m

(937)

These two equations can be solved to yield B and F in terms of A, B A F A

=

=

−1 i k h ¯2 m a V0

1 +

1

i k h ¯2 m a V0 2 + mi ka h¯V0

(938)

Thus, we again have | F |2 | B |2 + 2 |A| | A |2 T + R

=

1

= 1

(939)

and E is a continuous variable, in the range E > 0. We note that for large E, R → 0 and T → 1, as the potential becomes ineffective in scattering particles with sufficiently high energies.

4.1.18

Bound States of a delta function Potential

We shall now consider the bound states of a delta function potential which must have E < 0 and as a result E = −

¯ 2 κ2 h 2m

(940)

where κ is a positive real number. Accordingly, the wave function in the two regions must have the forms given by     φκ (x) = A exp − κ x + B exp + κ x (941) when 0 > x and

 φκ (x) = F exp

 − κx

(942)

when x > 0, corresponding to the analytic continuation k → i κ. In order for the wave function to be normalizable at x → − ∞, one must have A = 0. The matching conditions then yield B = F 215

(943)

and −

¯2 h 2m



 − κF

+

¯2 h 2m



 + κB

= V0 a F

(944)

which yields the energy eigenvalue condition ¯2 h κ = V0 a m

(945)

Thus, the energy eigenvalue is given by ( V0 )2  E = − 

(946)

2 h ¯2 m a2

which has the dimensions of energy. This result could have been obtained directly by noting that the reflection and transmission coefficients have poles at ¯2 h i k = − V0 a m

(947)

which is satisfied for imaginary values of k. Furthermore, since A can be neglected in comparison with B and F for this imaginary value of k, and since k = i κ, the wave function decays away from the origin   φκ (x) = B exp + κ x (948) in the region 0 > x and  φκ (x) = B exp

 − κx

(949)

for x > 0. The amplitudes of the decaying wave functions are equal (B = F ), 2 as can be seen from eqn(938) by noting that at the pole mi ka h¯V0 = − 1. Thus, we have recovered the bound state energy and the form of the wave function by analytic continuation to negative energies. The normalization of the wave function yields the amplitude, up to an arbitrary phase, as √ (950) B = κ

Solution in the Momentum Representation

216

Bound state wave function

Ψ0(x)

1.5

1

0.5

0 -2

-1

0

x/a

1

2

Figure 48: The bound state wave function of an attractive delta function potential. Note that the wave function has a discontinuous first derivative at x = 0. The bound states of the attractive one-dimensional delta function potential can be found by transforming to the momentum representation. The energy eigenvalue equation is   h2 ∂ 2 ¯ − − V0 a δ(x) ψ0 (x) = E0 ψ0 (x) (951) 2 m ∂x2 which after performing the Fourier Transform  1 Z ∞   1 2 φ0 (k) = dx ψ0 (x) exp − i k x 2π −∞

(952)

becomes

¯ 2 k2 h V0 a ψ0 (0) = E0 φ0 (k) φ0 (k) − √ 2m 2π On expressing the energy eigenvalue E0 in terms of κ E0 = −

¯ 2 κ2 h 2m

the equation is solved for φ0 (k) as   ψ0 (0) 1 2 m V0 a √ φ0 (k) = 2 2 + κ2 ) ( k ¯h 2π 217

(953)

(954)

(955)

The real space wave function ψ0 (x) is given by the inverse Fourier Transform of φ0 (k)  1 Z ∞   1 2 ψ0 (x) = dk φ0 (k) exp + i k x (956) 2π −∞ On performing the integral via Cauchy’s method, in which the contour is completed with a semi-circle in either the upper or lower half of the complex plane depending on the sign of x, one finds that the real space wave function is given by     m V0 a ψ0 (0) exp − κ | x | (957) ψ0 (x) = ¯h2 κ If the form of the solution for ψ0 (x) is consistent with the value of ψ0 (0) at x = 0, the value of κ must satisfy the equation κ =

m V0 a ¯h2

(958)

This condition determines the allowed value of the bound state energy. —————————————————————————————————— Example: The Bound States of a Triple delta function Potential. We shall examine the bound states of a particle of mass m moving in one dimension, in the presence of a potential V (x) given by   V (x) = − V0 a δ(x − a) + δ(x) + δ(x + a) (959) where V0 is a positive constant with units of potential and a is a length scale. The potential is an even function of x. The parity operator Pˆ is defined as the operator which, when acting on an arbitrary function ψ(x), has the effect Pˆ ψ(x) = ψ(−x)

(960)

Since the potential is an even function of x, the Hamiltonian commutes with the parity operator ˆ ] = 0 [ Pˆ , H (961) so that the parity is a conserved quantity. In other words, parity is a constant of motion. We shall search for bound states that are simultaneous eigenstates of the parity operator and the Hamiltonian. We shall look for eigenfunctions which are of the form   φn (x) = A exp

218

− κx

(962)

Triple delta function potential 5

V(x)/V0 a

0

-5

-10

-15

-20 -2

-1

0

1

2

x/a

Figure 49: The potential V (x) for three equally spaced attractive delta functions. in the region x > a and which are of the form     φn (x) = B exp − κ x + C exp + κ x

(963)

in the region a > x > 0. The forms of the solution for x < 0 are found from the above forms by requiring that the eigenfunctions either have even or odd parities. The above forms satisfy the energy eigenvalue equation in the regions where the potential is zero, as long as the energy eigenvalue of the bound state En is expressed in terms of κ via En = −

¯ 2 κ2 h 2m

(964)

The above form of the eigenfunction is a solution of the differential equation as long as it satisfies the differential equation at the boundary points x = 0 and x = ± a. The boundary conditions at x = a are the condition of continuity of the wave function and the condition that the discontinuity of its first derivative is related to the integral of the potential. These boundary conditions can be expressed as   A = B + C exp

219

2κa

(965)

and

 A

1 −

2 m V0 a ¯h2 κ



 = B − C exp

The above two equations can be combined to yield   m V0 a = B A 1 − ¯h2 κ and

m V0 a A = C exp h2 κ ¯



 2κa

(966)

(967)

 2κa

(968)

The eigenfunctions also have to satisfy boundary conditions at x = 0. The explicit forms of the boundary conditions at x = 0 depend on the parity of the eigenfunctions, therefore, we shall discuss the two cases separately. Odd Parity For odd parity solutions, φn (−x) = − φn (x), so the boundary condition at x = 0 reduces to φn (0) = 0 (969) or B = −C

(970)

Hence, one finds that B and C are completely determined by A, and A is determined, up to the phase, by the normalization condition. The above conditions imply that for odd parity solutions, κ must satisfy the equation     m V0 a2 m V0 a2 − κ a = exp − 2 κ a (971) h2 ¯ ¯h2 This equation can be solved graphically by plotting the left hand side and the right hand side as a function of κ a on the same graph, as seen in fig(50). The solutions for κ correspond to the points where the two curves cross. Both curves start at the same value m V0 a2 (972) ¯h2 when κ a = 0, and are both decreasing functions of κ a. The left hand side decreases linearly with increasing κ and becomes negative at κa =

m V0 a2 ¯h2

(973)

and the right hand side approaches the value of 0 exponentially as κ a → ∞. By considering the initial slopes of both curves, one finds that if 2

m V0 a2 > 1 ¯h2 220

(974)

Odd Parity Consistency Equation 2

(mV0a2/h2-κa) 1

(mV0a2/h2) exp [ -2κa ] 0

-1 0

1

2

3

κa

Figure 50: The graphical solution for the odd parity bound states, shown for two different strengths of the potential V0 . The left hand side of eqn(971) is shown by a red line, and the right hand side by a blue line. For large values of the potential or the separation, the equation (represented by the pair of solid lines) has one solution at a non-zero value of κ. For small values of V0 , the left hand and right hand sides of the equation are depicted by the dashed lines. The dashed curves only intersect at κ = 0. then the two curves also cross at a finite value of κ a. The value of κ a at the crossing yields a bound state solution with odd parity. The solution at κ a = 0 represents an odd parity zero energy resonance. Even Parity For even parity solutions, φn (−x) = φn (x). Therefore, the matching condition at the x = 0 boundary leads to     m V0 a B + C (975) B − C = ¯h2 κ or     m V0 a m V0 a = C 1 + (976) B 1 − ¯h2 κ ¯h2 κ On combining these conditions with the boundary conditions at x = a, one finds that κ a must satisfy the equation 2      m V0 a2 m V0 a2 m V0 a2 = κa + exp − 2 κ a (977) κa − h2 ¯ ¯h2 ¯h2 221

This equation determines the energy eigenvalues and energy eigenfunctions with even parity. The above transcendental equation can be solved graphically, as seen in fig(51). The left hand side and right hand side take on the same value

Even Parity Consistency Equations 4

(mV0a2/h2)(κa+mV0a2/h2) exp [ -2κa ]

3

(κa-mV0a2/h2)2 2

1

0 0

1

2

3

κa

Figure 51: The graphical solution of eqn(977) for the even parity bound states, shown for two different strengths of the potential. The left hand side of the equation is shown by a red line, and the right hand side by a blue line. For 2 large values of V0 h¯ma , the equation (represented by the solid lines) has two 2 solutions at non-zero values of κ. For small values of V0 , the equation only has one solution at a non-zero value of κ, located at the intersection of the dashed lines. at κ a = 0. The left hand side is always positive but touches zero at κa =

m V0 a2 ¯h2

(978)

and varies quadratically as κ a → ∞. The right hand side is monotonically decreasing and approaches zero as κ a → ∞. From consideration of the initial slope, it is found that there are three solutions of this equation if 2

m V0 a2 > 3 ¯h2

(979)

and, otherwise, only has two solutions. The solution at κ a = 0 corresponds to a zero energy bound state or resonance. The solutions at finite values of κ a have finite binding energies.

222

Bound State Energies

En m a2/h2

0

-1

n=1 n=2

-2

n=3

-3 0

0.5

1

1.5 2

mV0a /h

2

2

Figure 52: The bound state energy eigenvalues En , for the triple delta function potential.

Bound state wave functions mV0a2/h2 = 1.8

n=3 n=2

ψn(x)

0

n=1 0

0

-2

-1

0

1

2

x/a

Figure 53: The bound state wave functions φn (x), for the triple delta function potential.

223

The three bound state energies En are plotted as a function of κ a in fig(52). It is seen that the number of bound states increases as the value of V0 increases. As the separation between the wells is increased, all the bound state energies approach the asymptotic expression En → −

m V02 a2 2 ¯h2

(980)

expected for three isolated delta function potentials. The three bound state wave functions φn (x) are shown in fig(53). It is seen that the number of nodes increases with increasing energy. ——————————————————————————————————

4.1.19

Exercise 65

Find the reflection and transmission coefficients for a potential V (x) = V0 Θ(x) − V0 a δ(x)

(981)

where Θ(x) is the Heaviside step function, 1 for x > 0 Θ(x)

= 0 for x < 0

(982)

The potential is shown in fig(54). Show that the reflection and transmission coefficients add up to unity. ——————————————————————————————————

4.1.20

Solution 65

The energy eigenfunction corresponding to a scattering experiment where the incident beam is travelling towards the right is given by     φk (x) = A exp i k x + B exp − i k x x < 0   φk (x) = C exp i k 0 x x > 0 (983) where E =

h ¯ 2 k2 2 m

and E = V0 +

h ¯ 2 k02 2 m

The continuity condition leads to C = A + B 224

(984)

V(x) = -V0 a δ(x) + V0 Θ(x) 2

1

0 -0.5

0

V(x)

-1

0.5

1

-1

-2

-3

-4

x/a

Figure 54: The potential V (x) used in Exercise 65. whereas the discontinuity in the first derivative satisfies the equation   2 m V0 a i k0 + C = ik(A − B) ¯h2 Therefore

k + k0 − i 2k

2 m V0 a h ¯2



 k − k0 + i B = C 2k Thus, we have the transmission coefficient

2 m V0 a h ¯2



A = C



and

T

=

T

=

(985)

(986)

(987)

k 0 | C |2 k | A |2 (k +

k0

4 k k0 )2 + ( 2

m V0 a 2 ) h ¯2

(988)

and the reflection coefficient is given by R

=

R

=

| B |2 | A |2 ( k − k 0 )2 + ( 2 ( k + k 0 )2 + 225

m V0 a 2 ) h ¯2 2 m V0 a 2 ( h¯ 2 )

(989)

1

T(k)

0.8

T(k)

0.6

0.4

R(k)

0.2

0 1

1.5

2

2.5

3

k/k0

Figure 55: The transmission coefficient T (k) and the reflection coefficient R(k) calculated in Exercise 65.

——————————————————————————————————

4.1.21

Exercise 66

Find the reflection and transmission coefficients for a beam of particles incident on the potential V (x) given by   V (x) = − V0 a δ( x − a ) + δ( x + a ) (990) Also, find the bound states. ——————————————————————————————————

4.1.22

Solution 66

The energy eigenfunctions corresponding to the scattering states of energy E =

¯ 2 k2 h 2m

226

(991)

V(x) = - V0 a δ(x-a) - V0 a δ(x+a) 4

2

V(x)

-3

-2

E1

0

E0

-1

0

1

2

3

-2

-4

-6

-8

x/a

Figure 56: The potential for two attractive delta function potentials. can be expressed in terms of a steady state in which a particle beam of momentum ¯h k is incident from x → − ∞ or where a particle beam is incident from x → + ∞. We shall consider the solution corresponding to a beam incident from the left travelling to the right. The energy eigenfunction can be expressed as     φk (x) = A exp i k x + B exp − i k x x < −a     φk (x) = C exp i k x + D exp − i k x − a < x < a   φk (x) = F exp i k x a < x (992) where A is the amplitude of the incident beam and B and F are the amplitudes of the reflected and transmitted beams, respectively. The continuity conditions at x = − a is given by     A exp − i k a + B exp + i k a     = C exp − i k a + D exp + i k a

227

(993)

whereas, from integrating the energy eigenvalue equation, one finds the discontinuity in the first derivative satisfies      A exp − i k a − B exp + i k a      = C exp − i k a − D exp + i k a      2 m V0 a + C exp − i k a + D exp + i k a ik¯ h2 (994) This pair of equations can be solved for A and B in terms of C and D.      m V0 a m V0 a A = C 1 + + D exp + 2 i k a i k ¯h2 i k ¯h2      m V0 a m V0 a B = D 1 − − C exp − 2 i k a i k ¯h2 i k ¯h2 (995)

The boundary conditions at x = a are similar, the continuity equation is       C exp + i k a + D exp − i k a = F exp + i k a (996) whereas, from integrating the energy eigenvalue equation, one finds the discontinuity in the first derivative satisfies      C exp + i k a − D exp − i k a    2 m V0 a = F exp + i k a 1 + (997) i k ¯h2 This pair of equations can be solved to yield   m V0 a C = F 1 + i k ¯h2     m V0 a F exp + 2 i k a D = − i k ¯h2

(998)

The expressions for C and D can be substituted in the expressions for A and B, yielding the amplitudes of the reflected and transmitted beam 2  2    m V0 a A m V0 a − exp + 4 i k a = 1 + F ik¯ h2 i k ¯h2 228

B F

 = −

m V0 a ik¯ h2



 exp

 + 2ika

 + exp

 − 2ika (999)

The transmission coefficient is given by the inverse of the squared modulus of the first equation and the reflection coefficient is found by dividing the one equation by the other and then multiplying be its complex conjugate.

Self Consistency Equation 1

( m V0 a /h2)2 exp[ - 4 κ a]

0.8

0.6

( m V0 a /h2 - κ )2

0.4

0.2

0 0

0.2

0.4

0.6

0.8

1

κa

1.2

1.4

1.6

1.8

2

Figure 57: A graphical solution for the bound state energies. The bound states are given by the analytic continuation k → i κ and the values of κ are determined from the poles of the transmission and reflection coefficients, i.e. A = 0. Hence, we find  2  2   m V0 a m V0 a = exp − 4 κ a (1000) κ − h2 ¯ ¯h2 For large κ a the solution reduces to the solution for two independent delta function potentials, but these are subject to the exponentially small bonding and anti-bonding splitting     m V0 a2 m V0 a 1 ∓ exp − 2 (1001) κ ≈ ¯h2 ¯h2 For small κ a the splitting becomes larger and one obtains the values of κ as   m V0 a κ ≈ 2 ¯h2 229

Bound state wave functions 1

Ψ0(x)

Ψn(x)

0.5

0 -4

-3

-2

-1

Ψ1(x)

0

1

2

3

4

-0.5

-1

x/a

Figure 58: The bound state wave functions for the double delta function potential. κ

≈ 0

(1002)

This potential may serve as a one-dimensional model for a molecule, in which the levels are doubly degenerate. If the atoms are far apart, the energies just correspond to the energies of the individual atoms. However, when the atoms are brought closer together, there is an energy gain for doubly occupying the lowest energy state. This gain occurs through the bonding - anti-bonding splitting and results in the molecule being stabilized. Alternate Derivation of the Bound State Solutions. We shall find the bound states of particles moving in one dimension, in the presence of the potential V (x) given by   V (x) = − V0 a δ( x − a ) + δ( x + a ) (1003) The energy eigenvalue equation   h2 ∂ 2 ¯ − − V (x) ψα (x) = Eα ψα (x) 2 m ∂x2

230

(1004)

will be solved in the momentum space representation. On Fourier Transforming  φα (k) =

1 2π

 12 Z



 dx ψα (x) exp

 − ikx

(1005)

−∞

the energy eigenvalue equation, one obtains      h2 k 2 ¯ V0 a ψα (a) exp − i k a + ψα (−a) exp + i k a = Eα φα (k) φα (k) − √ 2m 2π (1006) On expressing the energy eigenvalue Eα in terms of κ Eα = −

¯ 2 κ2 h 2m

(1007)

the momentum space form of the energy eigenvalue equation is solved for φα (k), yielding        1 2 m V0 a 1 φα (k) = √ ψ (a) exp − i k a + ψ (−a) exp + i k a α α ( k 2 + κ2 ) ¯h2 2π (1008) The real space wave function ψα (x) is given by the inverse Fourier Transform of φα (k)  1 Z ∞   1 2 ψα (x) = dk φα (k) exp + i k x (1009) 2π −∞ The integrals are performed using Cauchy’s method, in which the contours are completed with semi-circles at infinity. One finds that the real space wave function is given by       m V0 a ψα (x) = ψ (a) exp − κ | x − a | + ψ (−a) exp − κ | x + a | α α h2 κ ¯ (1010) This solution must be consistent at the two points x = ± a. This yields the two consistency conditions        m V0 a m V0 a ψα (±a) 1 − = ψα (∓a) exp − 2 κ a (1011) h2 κ ¯ ¯h2 κ These equations determine the allowed values of κ. On combining these equations, one has    2  2   m V0 a m V0 a = exp − 4 κ a (1012) 1 − h2 κ ¯ ¯h2 κ which has two non-trivial solutions if   m V0 a2 > 1 2 ¯h2 231

(1013)

Otherwise, the above equation only has one non-trivial solution, and the trivial solution κ = 0. ——————————————————————————————————

4.1.23

Exercise 67

A particle of mass m moves in one dimension under the influence of an attractive delta function potential centered at the origin, and of strength V0 a. The particle is in the bound state at t = 0 and then the strength of the attractive potential is suddenly changed to V00 a. What is the probability that the particle will remain bound to the potential? ——————————————————————————————————

4.1.24

Solution 67

ˆ is governed by the The energy eigenstate Ψ0 (x) of the initial Hamiltonian H eigenvalue equation −

¯ 2 ∂2 h Ψ0 (x) − V0 a δ(x) Ψ0 (x) = E0 Ψ0 (x) 2 m ∂x2

(1014)

For x 6= 0 the eigenvalue equation reduces to −

¯ 2 ∂2 h Ψ0 (x) = E0 Ψ0 (x) 2 m ∂x2

(1015)

and has a solution of the form  Ψ0 (x)

= A exp

Ψ0 (x)

= B exp

 − κx



x > 0 

+ κx

x < 0

(1016)

where

¯2 2 h κ 2m The matching condition at x = 0 yields E0 = −

(1017)

Ψ0 ( +  ) = Ψ0 ( −  )

(1018)

A = B

(1019)

or

232

On integrating the differential equation over the infinitesimal interval between ( −  , +  ), one obtains   Z + Z + h2 ∂ 2 ¯ Ψ (x) − V a δ(x) Ψ (x) = E dx Ψ0 (x) dx − 0 0 0 0 2 m ∂x2 − − + h2 ∂ ¯ − Ψ0 (x) − V0 a Ψ0 (0) = E0 2  Ψ0 (0) 2 m ∂x −

(1020) or in the limit  → 0 +

¯2 h κ A − V0 a A = 0 m

(1021)

Hence, κ =

m a V0 ¯h2

(1022)

thus we have

m a2 V0 2 ¯h2 and the bound state wave function is given by   m a V0 Ψ0 (x) = A exp − | x | ¯h2 E0 = −

(1023)

(1024)

The magnitude of the constant A is given by the normalization condition Z ∞ dx | Ψ0 (x) |2 = 1 −∞   Z ∞ m a V0 2 |A| dx exp − 2 |x| = 1 ¯h2 −∞   Z ∞ m a V0 2 | A |2 dx exp − 2 x = 1 ¯h2 0 ¯h2 | A |2 = 1 (1025) m a V0 Hence, r

m a V0 (1026) ¯h2 up to an arbitrary phase. The initial ground state wave function is given by r   m a V0 m a V0 Ψ0 (x) = exp − | x | (1027) ¯h2 ¯h2 A =

233

The initial state wave function Ψ0 (x) can be expressed as a linear superpoˆ 0 , φn (x), via sition of the eigenstates of the final Hamiltonian, H X Ψ0 (x) = Cn φn (x) (1028) n

where the expansion coefficients are given by Z ∞ Cn = dx φ∗n (x) Ψ0 (x)

(1029)

−∞

ˆ 0 will result in a value for the energy The probability that a measurement of H 0 of En is given by P (n) = | Cn |2 (1030) The overlap between the initial bound state and the final bound state of potential V00 , is given by p   Z ∞ Z ∞ V0 V00 ma m a ( V0 + V00 ) ∗ dx exp − x dx φ0 (x) Ψ0 (x) = 2 ¯h2 ¯h2 0 −∞ p 2 V0 V00 (1031) = V0 + V00 The probability that the particle remains in the bound state, P (0), is given by 2 Z ∞ dx φ∗0 (x) Ψ0 (x) P (0) = −∞

=

4 V0 V00 ( V0 + V00 )2

(1032)

Thus, the probability that the particle ends up in an excited state is given by X P (n) Pexc = n6=0

= 1 − P (0) ( V0 − V00 )2 = ( V0 + V00 )2

(1033)

since the probabilities are normalized to unity. Note that if V0 = V00 no transitions take place. ——————————————————————————————————

4.1.25

Exercise 68

A particle moves in one dimension in a potential of the form V (x) V (x)

= ∞ for x < 0 = − V0 a δ( x − a ) 234

for x > 0

(1034)

15

V(x) = - V0 a δ(x-a) + V1 Θ(-x) δ

10

V(x)

5

0 -1

0

1

2

-5

-10

-15

x/a

Figure 59: The potential V (x) used in Exercise 68. The limit V1 → ∞ should be taken. The potential is shown in fig(59). Find the condition that determines the bound states. What is the minimum value of V0 for which a bound state appears? ——————————————————————————————————

4.1.26

Solution 68

We need to consider the bound state solutions of the energy eigenvalue equation   h2 ∂ 2 ¯ − + V (x) φ0 (x) = E0 φ0 (x) (1035) 2 m ∂x2

In the region when x > a the bound state solution must have the form of a decaying exponential   φ0 (x) = C exp − κ x (1036) where the bound state energy is E = − 235

¯ 2 κ2 h 2m

(1037)

Whereas in the region where 0 < x < a the wave function must have the form     φ0 (x) = A exp − κ x + B exp + κ x (1038) and vanish for x < 0. To satisfy the boundary condition at x = 0 one must have A = − B. The continuity condition at x = a yields        A exp − κ a − exp + κ a = C exp − κ a

(1039)

and on integrating the energy eigenvalue equation in an infinitesimal region about x = a to obtain the discontinuity of the first derivative in terms of the strength of the delta function potential,        − κ C exp − κ a − κ A exp − κ a + exp + κ a   2 m V0 a = − C exp − κ a (1040) h2 ¯

These two equations can be solved to yield   2 m V0 a coth κ a = − 1 − ¯h2 κ

(1041)

The bound state only just exists if κ → 0 in which case the above equation can be expanded in powers of κ. The solution at κ → 0 only exists if 2 m V0 a2 = 1 ¯h2

(1042)

——————————————————————————————————

236

4.2

The One-Dimensional Harmonic Oscillator

We shall find all the energy eigenfunctions of the one-dimensional Harmonic oscillator in a systematic way. The Hamiltonian of the one-dimensional harmonic oscillator can be written as   m ω 2 x2 ¯h2 ∂ 2 ˆ = + (1043) H − 2 m ∂x2 2

The Harmonic Oscillator Potential 5

E3 E2

V(x)

V(x) = 1/2 m ω2 x2

4 3

E1 E>0

2

E0

1 0 -1

-3

-2

-1

0

1

2

3

x

Figure 60: The Harmonic Oscillator Potential V (x) =

m ω2 2

x2 .

The Hamiltonian can be re-written in terms of the dimensionless variable r mω ξ = x (1044) ¯h Then, the Hamiltonian has the form   ¯h ω ∂2 2 ˆ H = − + ξ 2 ∂ξ 2

(1045)

which shows that h ¯ ω provides the natural energy scale for the quantum system. The Hamiltonian operator can be expressed in terms of its classical factorization and a constant either as     hω ¯ ∂ ∂ ˆ H = − + ξ + + ξ + 1 (1046) 2 ∂ξ ∂ξ 237

or as hω ˆ = ¯ H 2

 +

∂ + ξ ∂ξ

 −

∂ + ξ ∂ξ



 − 1

(1047)

since the factors do not commute. In fact, the commutator of the factors is given by     ∂ ∂ + + ξ , − + ξ = 2 (1048) ∂ξ ∂ξ 4.2.1

The Raising and Lowering Operators

Let us denote the factors by the operators a ˆ† and a ˆ since √ these factors are Hermitean conjugates. We shall also introduce a factor of 2 into their definition   1 ∂ a ˆ = √ + + ξ ∂ξ 2   1 ∂ a ˆ† = √ − + ξ (1049) ∂ξ 2 to simplify the commutation relations. Thus, these operators have the commutation relations [a ˆ, a ˆ† ] = + 1 (1050)

4.2.2

The Effect of the Lowering Operator

Let us consider the energy eigenvalue equation ˆ Ψn (x) = En Ψn (x) H

(1051)

with wave function Ψn (x) and energy eigenvalue En . The Hamiltonian when expressed in terms of the new operators is just   ˆ = ¯hω 2a ˆ† a ˆ + 1 H 2   h ¯ ω ˆ = H 2a ˆa ˆ† − 1 (1052) 2 The effect of the operators a ˆ on the eigenvalue equation can be found as ˆ Ψn (x) = En a a ˆH ˆ Ψn (x) (1053) but since the commutator of a ˆ with the Hamiltonian is   ¯h ω † ˆ a ˆH = a ˆ 2a ˆ a ˆ + 1 2   ¯h ω † = 2a ˆ a ˆ + 3 a ˆ 2   ˆ + ¯h ω a = H ˆ

238

(1054)

one finds eqn(1053) can be re-written as   ˆ + ¯h ω a H ˆ Ψn (x) = En a ˆ Ψn (x)

(1055)

Thus, on rewriting this we find ˆ a H ˆ Ψn (x) =



 En − ¯h ω

a ˆ Ψn (x)

(1056)

Hence, the wave function a ˆ Ψn (x)

(1057)

is an eigenfunction of the Hamiltonian with an eigenvalue of En − ¯h ω. Thus, the operator a ˆ when acting on an energy eigenfunction produces another energy eigenfunction with a lower energy eigenvalue. In other words, a ˆ Ψn (x) = Cn Ψn−1 (x)

(1058)

where Cn is a constant of proportionality. This property of a ˆ justifies naming it as the lowering operator.

4.2.3

The Ground State

The lowest energy state has eigenvalue E0 and, therefore, as there is no lower energy eigenfunction, one must have a ˆ Ψ0 (x) = 0

(1059)

This leads to the equation for the ground state wave function   ∂ + ξ Ψ0 (x) = 0 + ∂ξ This has the solution

 Ψ0 (x) = C exp

ξ2 − 2

(1060)

 (1061)

where C is the normalization constant that still has to be determined. In terms of the original variables, the ground state wave function is found as   m ω x2 Ψ0 (x) = C exp − (1062) 2 ¯h The normalization condition yields  |C | =

239

mω ¯h π

 14 (1063)

Since the lowest energy state satisfies eqn(1059) and due to the first form of the Hamiltonian given in eqn(1052), we deduce that the lowest energy eigenvalue is h ¯ ω 2 . The method implies that there exists higher-energy eigenstates that have energy eigenvalues which are larger by multiples of ¯h ω, that is   1 En = h ¯ ω n + (1064) 2

4.2.4

The Effect of The Raising Operator

The effect of the Hermitean conjugate of a ˆ can be found by considering its effect on the energy eigenvalue equation ˆ Ψn (x) = a a ˆ† H ˆ† En Ψn (x)

(1065)

On using the commutation relations, one finds   ¯h ω ˆ = a a ˆ† H ˆ† 2a ˆ† a ˆ + 1 2   ¯h ω = 2a ˆ† a ˆ − 1 a ˆ† 2   ˆ − ¯h ω a = H ˆ†

(1066)

so one finds that eqn(1065) becomes   ˆ H − ¯h ω a ˆ† Ψn (x) = En a ˆ† Ψn (x)

(1067)

Thus, on re-writing the above equation, we find   † ˆ H a ˆ Ψn (x) = En + ¯h ω a ˆ† Ψn (x)

(1068)

Hence, the wave function a ˆ† Ψn (x)

(1069)

is an eigenfunction of the Hamiltonian with an eigenvalue of En + h ¯ ω, i.e. En+1 . Thus, the operator a ˆ† when acting on an energy eigenfunction produces another energy eigenfunction with a higher-energy eigenvalue. ∗ a ˆ† Ψn (x) = Cn+1 Ψn+1 (x)

(1070)

∗ is a constant of proportionality related to that in eqn(1058). This where Cn+1 property of a ˆ† justifies naming it as the raising operator.

240

4.2.5

The Normalization

The constant of proportionality for the raising operator is related to the constant of proportionality for the lowering operator as can be found by considering the definition of the Hermitean conjugate of an operator  Z +∞ ∗ Z +∞ ∗ ∗ † dx Ψn−1 (x) a ˆ Ψn (x) = dx Ψn (x) a ˆ Ψn−1 (x) (1071) − ∞

− ∞

This implies that, with properly normalized eigenfunctions and on using the definition in eqn(1068) and its consequence eqn(1070), the coefficients of proportionality satisfy   ∗

Cn∗

Cn =

(1072)

as expected. The constant of proportionality can be found, up to an arbitrary phase, from the energy eigenvalue equation     1 hω ¯ † 2a ˆ a ˆ + 1 Ψn (x) = h ¯ ω n + Ψn (x) (1073) 2 2 Thus, on using the equations for the raising and lowering operators together with the constants of proportionality, one finds | Cn |2 = n Hence, on choosing a real phase for the Cn , we have the equations √ n Ψn−1 (x) a ˆ Ψn (x) = √ † a ˆ Ψn (x) = n + 1 Ψn+1 (x)

4.2.6

(1074)

(1075)

The Excited States

All the energy eigenfunctions can be obtained from the ground state wave function by successive action of the raising operator, Hence, we have = a ˆ† Ψ0 (x) r r   1 h ¯ ∂ mω Ψ1 (x) = √ − + x Ψ0 (x) m ω ∂x ¯h 2 r r   14     mω ¯h ∂ mω m ω x2 Ψ1 (x) = − + x exp − 4 ¯h π m ω ∂x ¯h 2 ¯h (1076) Ψ1 (x)

241

The higher-energy eigenfunctions are given by iteration Ψn+1 (x)

=

Ψn+1 (x)

=

a ˆ† p Ψn (x) (n + 1) r r   ¯ h ∂ mω 1 p − + x Ψn (x) m ω ∂x ¯h 2(n + 1) (1077)

Then the n-th excited state wave function is given in terms of the ground state wave function Ψ0 (x) by acting on it by the raising operator n times Ψn (x)

= = =

(a ˆ† )n √ Ψ0 (x) n! 1    m ω x2 (a ˆ† )n mω 4 √ exp − hπ ¯ 2 ¯h n! r r  n  1   1 ¯h ∂ mω mω 4 m ω x2 √ − + x exp − m ω ∂x ¯h ¯h π 2 ¯h 2n n! (1078)

Furthermore, using the operator identity r r   h ¯ ∂ mω − + x ≡ m ω ∂x ¯h r     ¯h ∂ m ω x2 m ω x2 exp − ≡ ( − 1 ) exp + m ω ∂x 2 ¯h 2 ¯h (1079) one finds the eigenstates are given by the expression n    r ¯h ∂ m ω x2 ( − 1 )n m ω x2 exp − Ψn (x) = √ exp + m ω ∂x 2 ¯h 2 ¯h 2n n!   14   mω m ω x2 × exp − ¯h π 2 ¯h (1080) The solution is recognized as involving the n-th order Hermite polynomial Hn (x) given by r   r n   mω m ω x2 ¯h ∂ m ω x2 n x = ( − 1 ) exp + exp − Hn h ¯ ¯h m ω ∂x ¯h (1081) Thus, we have found the expression for the general energy eigenfunction for the one-dimensional harmonic oscillator as   1  r  1 mω 4 m ω x2 mω √ Ψn (x) = exp − Hn x (1082) hπ ¯ 2 ¯h ¯h 2n n! 242

These states are all properly normalized to unity, as the ground state wave function was properly normalized.

Harmonic Oscillator wave functions Ψ7(x) Ψ6(x)

Ψn(x)

Ψ5(x) Ψ4(x) Ψ3(x) Ψ2(x) Ψ1(x) -4

-3

-2

-1

0

1

2

3

4

x

Figure 61: The ground state Ψ0 (x) and excited state wave functions Ψn (x) of the one-dimensional harmonic oscillator. The quantum probability density for finding a particle with energy E at position x should be compared with the classical probability density if the initial position is unknown. Unlike the classical probability density, the quantum probability density shows spatial oscillations. If the spatial oscillations are averaged over, then it is seen that as the energy increases, the quantum probability density approaches the classical density. ——————————————————————————————————

4.2.7

Exercise 69

Find an expression for the expectation value of x4 , in the n-th excited state of the harmonic oscillator. ——————————————————————————————————

243

0.8

Pn(x) = l Ψn(x) l 2

0.6

n=4 0.4

0.2

0 -4.5

-3.5

-2.5

-1.5

-0.5

0.5

1.5

2.5

3.5

4.5

x

Figure 62: Comparison of the quantum mechanical probability densities for finding a particle at position x, P (x) = | Ψn (x) |2 with the classical probability densities for n = 4. 4.2.8

Solution 69

The average value of x4 in the n-th energy eigenstate of the harmonic oscillator can be found be expressing x in terms of the raising and lowering operators r ¯h x = ( a† + a ) (1083) 2mω Hence x4

Z

+∞

dx φ∗n (x) x4 φn (x)

= −∞

 =

¯ h 2mω

2 Z

+∞

dx φ∗n (x) ( a† + a )4 φn (x)

−∞

(1084) The only non-zero terms are those involving the same number of raising and lowering operators as terms where these numbers are different can be related to the matrix elements of energy states with different values of n which, because of the orthogonality of non-degenerate energy eigenstates, are zero. The non-zero

244

0.8

n=8 Pn(x) = l Ψn(x) l 2

0.6

0.4

0.2

0 -4.5

-3.5

-2.5

-1.5

-0.5

0.5

1.5

2.5

3.5

4.5

x

Figure 63: Comparison of the quantum mechanical probability densities for finding a particle at position x, P (x) = | Ψn (x) |2 with the classical probability densities for n = 8. terms are x4

=



¯ h 2mω

2 Z



¯ h 2mω

2 

+∞

dx φ∗n (x) −∞  † † + a a a a + H.c. φn (x)

=



a† a† a a + a† a a† a

n(n − 1) + nn + n(n + 1)

 +(n + 1)n + (n + 1)(n + 1) + (n + 2)(n + 1)  x4

=

¯ h 2mω

2

3 ( 2 n2 + 2 n + 1 )

(1085)

——————————————————————————————————

245

4.2.9

Exercise 70

Show that for a harmonic oscillator in the n-th energy eigenstates, the uncertainty in the position and momentum satisfy the equation 2n + 1 ∆xrms ∆prms = h ¯ (1086) 2

——————————————————————————————————

4.2.10

Solution 70

The expectation value of the ∆x2 is equal to the expectation value of x2 Z ¯h 2 ∆x = dx φ∗n (x) ( a† + a )2 φn (x) 2mω ¯h (2n + 1) = 2mω In the first line we have used the fact that the term Z dx φ∗n (x) ( a† )2 φn (x) (1087) and the Hermitean conjugate term both vanish identically. In the second line, we have used the fact that the state φn (x) is an eigenstate of the number operator a ˆ† a ˆ = n ˆ with eigenvalue n. Likewise Z mω ∆ˆ p2 = − ¯h2 dx φ∗n (x) ( a† − a )2 φn (x) 2 ¯h hmω ¯ (2n + 1) = 2 Combining these, we find the equality ∆prms ∆xrms =

¯ h (2n + 1) 2

(1088)

——————————————————————————————————

4.2.11

Exercise 71

A particle is moving in one dimension in a potential V (x) which is given by m ω2 2 x x > 0 2 V (x) → ∞ x < 0 V (x)

=

Find the energy eigenfunctions and eigenvalues.

246

(1089)

4.2.12

Time Development of the Harmonic Oscillator

Given an initial condition Ψ(x, 0), one can expand this initial wave function in terms of the energy eigenstates φn (x) of the harmonic oscillator via X Ψ(x; 0) = Cn φn (x) (1090) n

where the expansion coefficients may be found from Z +∞ Cn = dx φ∗n (x) Ψ(x; 0)

(1091)

− ∞

The wave function at future times t can be found from the solution of the Schr¨ odinger equation, and as the Hamiltonian is time independent the solution can be expressed as  Ψ(x; t) = exp

ˆ t  H Ψ(x; 0) − i ¯h

(1092)

which gives rise to the expression for the time-dependent wave function as Ψ(x; t)

=

X

Cn

n

Ψ(x; t)

=

X

ˆ t  H φn (x) exp − i ¯h   En t exp − i φn (x) ¯h 

Cn

n

(1093)

since φn (x) satisfy the energy eigenvalue equation ˆ φn (x) = En φn (x) H

(1094)

The motion of the particle can be viewed through examination of the average position x(t) or momentum p(t). The average position is given by the expectation value Z +∞ x(t) = dx Ψ∗ (x; t) x Ψ(x; t)

(1095)

− ∞

which, on using the expression in eqn(1093), leads to   Z +∞ X ( Em − E n ) t ∗ x(t) = Cm Cn exp + i dx φ∗m (x) x φn (x) h ¯ − ∞ n,m (1096)

247

The matrix elements of x between the two energy eigenfunctions can be evaluated by expressing the x in terms of the raising and lowering operators, r   ¯h 1 ∂ ∂ x = ξ + + ξ − mω 2 ∂ξ ∂ξ r   ¯h = a + a† (1097) 2mω Then the matrix elements can be found to be r   Z +∞ √ √ ¯h ∗ dx φm (x) x φn (x) = n + 1 δm,n+1 + n δm,n−1 2mω −∞ (1098) On substituting this expression back into eqn(1096) and performing the double summation and then using En+1 − En = h ¯ ω, one finds Z +∞ x(t) = dx Ψ∗ (x; t) x Ψ(x; t) −∞ r   X √ h ¯ ∗ = n + 1 Cn+1 Cn exp + i ω t + 2mω n   √ ∗ Cn exp − i ω t + n Cn−1 r    X √ h ¯ ∗ n+1 Cn+1 Cn exp + i ω t + = 2mω n   ∗ + Cn Cn+1 exp − i ω t (1099) where we have shifted the summation index by one in the second term. Since the first term is equal to the complex conjugate of the second term, the expectation value is real. This is as it must be, because x is a Hermitean operator. Let us denote the expansion coefficients in terms of an amplitude and phase   Cn = | Cn | exp i ϕn (1100) then we have r   2 ¯h X √ n + 1 | Cn+1 | | Cn | cos ω t + ϕn − ϕn+1 x(t) = mω n (1101) Hence, the position oscillates with frequency ω, just like the classical value. However, if the phase differences ϕn+1 − ϕn are randomly distributed, destructive interference may mask the oscillations.

248

One can evaluate the expectation value of the momentum via a similar procedure. The expectation value is given by Z +∞ ∂ p(t) = − i ¯h dx Ψ∗ (x; t) Ψ(x; t) (1102) ∂x − ∞ which yields p(t) = − i ¯ h

X n,m

∗ Cm



( E m − En ) t Cn exp + i ¯h

Z

+ ∞

dx φ∗m (x)

− ∞

∂ φn (x) ∂x

(1103) The matrix elements can be evaluated by expressing the derivative in terms of the raising and lowering operators r   ∂ mω 1 ∂ ∂ = ξ + − ξ + ∂x ¯h 2 ∂ξ ∂ξ r   mω a − a† (1104) = 2 ¯h Hence, we have the matrix elements r   Z +∞ √ √ ∂ mω φn (x) = dx φ∗m (x) n δm,n−1 − n + 1 δm,n+1 ∂x 2¯ h − ∞ (1105) On substituting the above matrix elements into the expectation value of the momentum, one finds r √ 2mω X p(t) = − ¯ h | Cn+1 | | Cn | n + 1 sin ( ω t + ϕn − ϕn+1 ) ¯h n (1106) Thus, we see that the average momentum oscillates out of phase with the average position, and ∂ p(t) = m x(t) (1107) ∂t which is just the same as the classical mechanical expression. ——————————————————————————————————

4.2.13

Exercise 72

Consider a harmonic oscillator in a state given by Ψ(x) = √

1 2s

n=N X+s

 exp

n=N −s

249

 − inϕ

φn (x)

(1108)

where φn (x) is the n-th excited state of the harmonic oscillator and N  s  1. Find the time dependence of x(t) and p(t). Compare these expectation values with the classical expressions. ——————————————————————————————————

4.2.14

Solution 72

This type of state is a coherent state as it is a superposition of a very large number of energy eigenstates, where the exists a definite phase relation between the component states. The time dependence of the coherent states is given by     n=N X+s 1 Ψ(x, t) = √ exp − i n ϕ exp − i n ω t φn (x) (1109) 2 s n=N −s As a result, the component states always have a simple phase relationship between them. The expectation value of x ˆ is given by  Z ∞ X 1 0 x(t) = exp i ( n − n ) ( ϕ + ω t ) dx φ∗n0 (x) x φn (x) (1110) 2s −∞ 0 n,n

However, the operator x ˆ can be expressed in terms of the creation and annihilation operators via r ¯h x ˆ = ( a† + a ) (1111) 2mω Then, the matrix elements of x ˆ between the energy eigenstates φn (x) and φ∗n (x) can be evaluated as r   Z ∞ √ √ ¯h ∗ 0 0 dx φn0 (x) x φn (x) = δn ,n+1 n + 1 + δn ,n−1 n 2mω −∞ (1112) Thus, the expectation value is calculated as r   X  √ 1 h ¯ x(t) = n + 1 exp i ( ϕ + ω t ) δn0 ,n+1 2s 2mω n,n0    √ + n exp − i ( ϕ + ω t ) δn0 ,n−1 (1113) If one only retains the terms of leading order in N1 and 1s , one finds that the expectation value of the position is given by r ¯h N x(t) = cos( ω t + ϕ ) (1114) 2mω 250

Likewise, for the expectation values of the momentum, one can show that r mωN p(t) = − sin( ω t + ϕ ) (1115) 2 ¯h The expectation value of the momentum is related to the expectation value of the position via the classical relation pˆ(t) = m

∂x(t) ∂t

Schr¨ odinger12 determined that the coherent state   n   X − i n ϕ λ exp ∞ 1 2 √ Ψλ (x) = exp − λ φn (x) 2 n! n=0 is an exact eigenstate of the lowering operator. This can be seen as   a Ψλ (x) = λ exp − i ϕ Ψλ (x)

(1116)

(1117)

(1118)

Coherent states play an important role in laser optics as they approximate classical states, containing many quanta, that have definite phases. ——————————————————————————————————

4.2.15

Hermite Polynomials

The Hermite Polynomials Hn (z) can be defined as   n   n ∂ 2 Hn (z) = − 1 exp + z 2 exp − z ∂z n

(1119)

From this definition, we can see that they satisfy the differential equation ∂2 ∂ Hn − 2 z Hn + 2 n Hn (z) = 0 2 ∂z ∂z

(1120)

This is proved by evaluating the first and second derivatives of Hn (z). The first derivative is given by   n   n ∂ ∂ 2 2 Hn = 2 z − 1 exp + z exp − z ∂z ∂z n   n   n ∂ 2 2 − − 1 exp + z 2 z exp − z (1121) ∂z n 12 E.

Schr¨ odinger, Naturwissenschaften, 14, 664 (1926).

251

and the second derivative is given by ∂2 Hn ∂z 2

  ∂n 2 = (4z + 2) − 1 exp + z exp − z ∂z n     n n ∂ 2 2 −8z − 1 exp + z z exp − z ∂z n   n   n ∂ 2 2 2 + − 1 exp + z ( 4 z − 2 ) exp − z ∂z n (1122) 2

n



2



The terms + 2 and − 2 in the round brackets cancel. On forming the combination ∂2 ∂ Hn − 2 z Hn (1123) ∂z 2 ∂z we have   n   n ∂ ∂ ∂2 2 2 Hn − 2 z Hn = − 4 z − 1 exp + z z exp − z ∂z 2 ∂z ∂z n     n n ∂ 2 2 2 + − 1 exp + z 4 z exp − z ∂z n (1124) The righthand side can be simplified as     n+1 n ∂ 2 exp − z = +2z − 1 exp + z 2 ∂z n+1     n n ∂ ∂ 2 − − 1 exp + z 2 2 z exp − z ∂z n ∂z (1125) which after commuting the z in the last term to the front, we have   n   n ∂ 2 2 = −2n − 1 exp + z exp − z ∂z n = − 2 n Hn (z)

(1126)

Thus, we have found that the Hermite polynomials satisfy the equation ∂2 ∂ Hn − 2 z Hn + 2 n Hn (z) = 0 ∂z 2 ∂z

(1127)

as was to be proved. This differential equation also has an integral representation of its solution. The integral representation of the solution is given by   n Z +∞ 2n Hn (z) = √ du exp − u2 z + iu (1128) π − ∞ 252

This can be shown by first evaluating the terms   n−2 Z +∞ ∂2 2n 2 √ du exp − u z + i u H (z) = n ( n − 1 ) n ∂z 2 π − ∞   n−1 Z +∞ n 2 ∂ 2 du exp − u z + iu (1129) Hn (z) = n √ ∂z π − ∞ and then by forming the expression ∂ Hn (z) 2 n Hn (z) − 2 z ∂z   n−1 Z + ∞ 2n 2 = in √ du 2 u exp − u z + iu π − ∞    n−1 Z +∞ 2n ∂ = −in √ du exp − u2 z + iu ∂u π − ∞ (1130) On integrating by parts, one finds this simplifies to  n−1   Z +∞ 2n ∂ 2 = in √ du exp − u z + iu ∂u π − ∞   n−2 Z +∞ 2n du exp − u2 z + iu = −n(n − 1) √ π − ∞ = −

∂2 Hn (z) ∂z 2

(1131)

Thus, the integral expression of eqn(1128) satisfies the differential equation ∂2 ∂ Hn − 2 z Hn + 2 n Hn (z) = 0 ∂z 2 ∂z

(1132)

and, therefore, is a representation of the Hermite polynomials. The integral representation can be used to yield the explicit forms of low-order Hermite polynomials. Expressions for first few lowest order Hermite polynomials Hn (z) are given in Table(2). We can show that ∂ Hn (z) = 2 n Hn−1 (z) ∂z

(1133)

This is seen by examining the first derivative   n   n ∂ ∂ 2 2 Hn = 2 z − 1 exp + z exp − z ∂z ∂z n   n   n ∂ 2 2 − − 1 exp + z 2 z exp − z (1134) ∂z n 253

Table 2: The Lowest Order Hermite Polynomials Hn (z). n

0 1 2 3 4 5

Hn (z)

H0 (z) H1 (z) H2 (z) H3 (z) H4 (z) H5 (z)

1 2z 4 z2 − 2 8 z 3 − 12 z 4 16 z − 48 z 2 + 12 32 z 5 − 160 z 3 + 120 z

On commuting the term proportional to z in the second term to the front, we have   n−1   n ∂ ∂ 2 Hn = − 2 n − 1 exp + z 2 exp − z ∂z ∂z n−1 = 2 n Hn−1 (z) (1135) Thus, the derivative of the n-th order Hermite polynomial is related to the Hermite polynomial of (n − 1)-th order. From this relationship, we can derive the generating function for the Hermite polynomials   ∞ X Hn (z) n 2 exp − t + 2 t z = t (1136) n! n = 0 The proof starts with examining F (z, t), which is defined via F (z, t) =

∞ X Hn (z) n t n! n = 0

(1137)

On taking the derivative with respect to z, one has ∂ F (z, t) ∂z

= = =

∞ X n = 0 ∞ X

∂ Hn (z) n t ∂z n! 2n

n = 0 ∞ X

2

n = 0

Hn−1 (z) n t (n − 1)!

= 2 t F (z, t)

254

Hn−1 (z) n t n!

(1138)

The equation ∂ F (z, t) = 2 t F (z, t) ∂z

(1139)

can be integrated to yield ln

F (z, t) = 2tz F (0, t)

(1140)

Thus, we have  F (z, t) = F (0, t) exp

 2zt

(1141)

∞ X Hn (0) n t n! n = 0

(1142)

We can evaluate F (0, t) from Hn (0) as F (0, t) =

We shall re-write the expression for Hn (z) in terms of a power series and then set z = 0,   n X n m ( z 2 )m ∂ Hn (z) = − 1 exp + z 2 − 1 (1143) n ∂z m! m The only term that remains on setting z = 0 is the term with n = 2 m. Hence, only the even order Hermite polynomials remain finite at z = 0. We find that these are given by H2m (0) =

− 1

m ( 2 m )! m!

(1144)

Thus, we have ∞ X Hn (0) n F (0, t) = t n! n = 0 m ∞ X − 1 = t2m m! m = 0   2 = exp − t

(1145)

Hence, we obtain  F (z, t)

+ 2 z t − t2

=

exp

=

∞ X Hn (z) n t n! n = 0



(1146)

which is the generating function expansion for the Hermite polynomials. ——————————————————————————————————

255

4.2.16

Exercise 73

The initial wave function of a particle of mass m in a harmonic potential of frequency ω is given by  Ψ(x, 0) =

mω π ¯h

 14

 exp



mω ( x − a )2 2 ¯h

 (1147)

Find the probability that a measurement of the energy will give the result En = h ¯ ω(n +

1 ) 2

(1148)

——————————————————————————————————

4.2.17

Solution 73

Using the generating function expansion, we can decompose the initial wave function as  1   mω 4 mω 2 Ψ(x, 0) = exp − (x − a) π¯ h 2 ¯h 1    mω 4 mω 2 = exp − x π¯ h 2 ¯h   mω 2 mω × exp − a + xa 2 ¯h ¯h  14    mω mω 2 = exp − x π¯ h 2 ¯h     mω 2 mω a mω 2 a exp − a + 2 x × exp − 4 ¯h 4 ¯h ¯h 2    14  mω 2 mω exp − = a π¯ h 4 ¯h p mω    n X x) m ω 2 Hn ( m ω a2 2 h ¯ × exp − x 2 ¯h n! 4 ¯h n n   X   mω 2 φn (x) m ω a2 2 √ = exp − a (1149) 4 ¯h 2 ¯h n! n where φn (x) is the n-th energy eigenfunction given by   m ω 2 r   1 exp − 2 h¯ x mω 4 mω φn (x) = H ( x) n 1 n π¯ h ¯h 2 2 ( n! ) 2 256

(1150)

Thus, as the probability P (n) is given by the square of the expansion coefficient, we have  n m ω a2 2 h ¯

P (n) =

 exp

n!

mω 2 − a 2 ¯h

 (1151)

——————————————————————————————————

4.2.18

Exercise 74

Find the time dependence of the state which has the initial wave function   1  mω 4 mω 2 (x − a) (1152) Ψ(x, 0) = exp − 2 ¯h π ¯h and, hence, find the time dependence of the probability density P (x, t) of finding the particle at position x. ——————————————————————————————————

4.2.19

Solution 74

From the previous example, we have the expansion of the initial wave function in terms of energy eigenstates,  1   mω 4 mω 2 (x − a) Ψ(x, 0) = exp − π ¯h 2 ¯h    n m ω 2 X φn (x) m ω a2 2 √ = exp − a (1153) 4 ¯h 2 ¯h n! n The time dependence of the wave function is then found from the time dependence of the energy eigenstates  ˆ t  H Ψ(x, t) = exp − i Ψ(x, 0) ¯h     ωt mω 2 = exp − i exp − a 2 4 ¯h    n X m ω a2 2 φn (x) √ × exp − i n ω t (1154) 2 ¯h n! n which can be re-summed to yield   1    ωt mω 4 mω 2 Ψ(x, t) = exp − i exp − x π ¯h 2 2 ¯h 257

 × exp



mω 2 a (1 + e 4 ¯h

− 2 i ω t

) +

mω xae ¯h

− i ω t



(1155) Hence, we find the probability density P (x, t) as P (x, t)

= | Ψ(x, t) |2  1   mω 2 mω 2 = exp − ( x − a cos ω t ) π¯ h ¯ h

(1156)

Hence we see that the state, which was initially displaced from the equilibrium state by a distance a, performs oscillations of amplitude a and frequency ω just like a classical particle in the same potential. ——————————————————————————————————

4.2.20

The Completeness Condition

As the Hamiltonian operator for the harmonic oscillator is a Hermitean operator, the energy eigenfunctions form a complete set. The completeness condition for the harmonic oscillator energy eigenfunctions can be proved by using the integral representation and the generating function. We take the generating function   F (x, t) = exp + 2 x t − t2 =

∞ X Hn (x) n t n! n = 0

(1157)

then let t = y + i u, so     exp + 2 x y − y 2 exp 2 i ( x − y ) u + u2 =

∞ X Hn (x) ( y + i u )n n! n = 0

 On multiplying this equation by exp

− u2

(1158)  and then integrating over u from

− ∞ to + ∞, we find     Z +∞ 2 du exp + 2 x y − y exp 2 i ( x − y ) u − ∞ ∞ X

=

n = 0

Hn (x) n!

Z

+ ∞

 du exp

− ∞

258

− u2



( y + i u )n

(1159)

so we have   √ π π exp + 2 x y − y 2 δ( x − y ) = n 2

∞ X Hn (x) Hn (y) (1160) n! n = 0

The above equation can be rearranged to yield δ( x − y ) =

∞ X n = 0

1 √ exp n 2 n! π

x2 − 2





 Hn (x) exp

y2 − 2

 Hn (y)

(1161) On identifying the normalized energy eigenfunctions φn (x) with the product of the exponential and the Hermite polynomials, one finds the completeness relation for the energy eigenfunctions of the Hermitean Hamilton operator X δ( x − y ) = φ∗n (x) φn (y) (1162) n

The completeness condition allows an arbitrary function to be expanded in terms of the energy eigenstates.

259

4.3

Dual-symmetry

Let us assume that we know the ground state solution of the energy eigenvalue equation for a particle moving in one dimension in the presence of a specific potential V (x). Let the ground state energy be denoted by E0 and the ground state wave function be denoted by φ0 (x). Using these definitions, the energy eigenvalue equation has the form 2

 −

2

∂ ¯ h 2m ∂x2

ˆ φ0 (x) = E0 φ0 (x) H  + V (x) φ0 (x) = E0 φ0 (x)

ˆ − E0 with an operator Alternatively, we can identify H    00 ¯h2 ∂2 φ0 (x) ˆ H − E0 = − + 2m ∂x2 φ0 (x)

(1163)

(1164)

This (energy-shifted) Hamiltonian of eqn(1164) can be factorized as the product of a pair of Hermitean conjugate operators. The first operator is defined by   0  ¯h ∂ φ0 (x) ˆ A = √ − (1165) φ0 (x) 2 m ∂x and its Hermitean conjugate is then found to be   0  ¯h ∂ φ0 (x) † ˆ √ A = − − ∂x φ0 (x) 2m

(1166)

The Hamiltonian (up to the definition of the ground state energy) is given by the product ˆ − E0 = Aˆ† Aˆ H (1167) The pair of Hermitean conjugate operators have a commutation relation given by  2  0 2   00 ¯h φ0 (x) φ0 (x) − (1168) [ Aˆ , Aˆ† ] = 2 2m φ0 (x) φ0 (x) For the case of the Harmonic Oscillator, Aˆ and Aˆ† respectively are proportional to the lowering and raising operators, and their commutator is merely the constant ¯h ω. ˆ d of the Hamiltonian H ˆ is given by the product of the The dual partner H pair of operators taken in the reverse order ˆ d = Aˆ Aˆ† H

(1169)

The dual potential Vd (x) is defined in terms of the dual Hamiltonian. The dual potential is found to be Vd (x)

= V (x) − E0 + [ Aˆ , Aˆ† ]   ¯h2 ∂ φ00 (x) = V (x) − E0 − m ∂x φ0 (x) 260

(1170)

The pair of potentials V (x) and Vd (x) are dual partner potentials. For the Harmonic Oscillator, the potential and the dual potential are the same except for a constant shift of the energy. In general, Vd (x) and V (x) have the same energy level spectrum, En − E0 = End0 . The exceptional case is the zero energy eigenvalue for the ground state of V (x), and this exception occurs since the operator Aˆ annihilates the ground state. The equality between the shifted eigenvalues is proved by noting that, if ˆ φn = (En − E0 ) φn H

(1171)

then ˆd H



Aˆ φn



  † ˆ ˆ ˆ = AA A φn   ˆ ˆ = A H φn   ˆ = (En − E0 ) A φn

(1172)

Hence, End0 = (En − E0 ), unless Aˆ φn = 0. Likewise, one can establish the ˆ d with the eigenstates of H. ˆ inverse relationship between the eigenstates of H ˆ s with eigenvalue End , then Aˆ† φdn (x) is an That is, if φdn (x) is an eigenstate of H ˆ with eigenvalue E d . Thus, Aˆ and Aˆ† connect states of different eigenstate of H n Hamiltonians that have the same energies. Duality and the Infinite Square Well The infinite square well has an infinite number of bound states with energy eigenvalues given by  2 ¯h2 πn En = (1173) 2m L and energy eigenfunctions are given by r   2 nπx φn (x) = sin L L

(1174)

where n is a positive integer, excluding zero. Since the logarithmic derivative of the ground state wave function is given by  0      φ1 (x) π πx = cot (1175) φ1 (x) L L the dual partner potential is found to be  2     ¯h2 π πx 2 Vd (x) = csc − 1 2m L L 261

(1176)

Eigenfunctions of the Infinite Square Well

0

n=3

φn(x)

n=2 0

n=1

0

0

0.2

0.4

0.6

0.8

1

x/L

Figure 64: The first three eigenfunctions φn (x) for the infinite square well potential. The dual energy eigenstates are given, up to an arbitrary multiplicative constant, by φdn (x) ∝ Aˆ φn (x)        ∂ π πx nπx − cot sin ∝ ∂x L L L   π x #     " sin n L nπx πx   ∝ n cos − cos L L π x L

where φd1 (x) and E1d are absent.

262

(1177)

Dual Potential

Vd(x) [ 2m/h2 (L/π)2 ]

10

Ed3

8

6

4

Ed2

2

0 0

0.2

0.4

0.6

0.8

1

x/L

Figure 65: The potential V d (x) dual to the infinite square well potential. The energy eigenvalues are marked by the dashed horizonatl lines.

4.4

Bargmann Potentials

Bargmann potentials are an interesting class of one-dimensional potentials. Bargmann potentials have the very unusual property that they are reflectionless potentials, which means that a plane wave incident on the potential does not produce a reflected wave. The potential produces bound states and does have an effect on the states with E > 0 in that the transmitted wave experiences a phase shift relative to the incident wave. We shall consider solutions of the family of energy eigenvalue equations governed by the integer parameter n. The eigenvalue equation is given by   ¯h2 ∂ 2 x 2 Ψα,n (x) − V0 n ( n + 1 ) sech Ψα,n (x) = Eα Ψα,n (x) − 2 m ∂x2 ξ (1178) where the length scale ξ is given by ξ2 =

¯2 h 2 m V0

263

(1179)

Dual wave functions

0

φ n(x)

n=3 n=2

0

0

0.2

0.4

x/L

0.6

0.8

1

Figure 66: The lowest two energy eigenstates φdn (x) of the dual potential. For n = 0, the Bargmann energy eigenvalue equation reduces to the free particle problem. The general solution of the free particle problem is simply     Ψα,0 (x) = A exp + i k x + B exp − i k x (1180) where the energy eigenvalue is given by Eα,0 =

¯ 2 k2 h 2m

(1181)

We can relate the general scattering solution for arbitrary n to the general solution for the free particle problem (where n = 0) by an iterative method. This relation involves dual-symmetry. We shall show that    ∂ x Ψα,n−1 (x) (1182) Ψα,n (x) = ξ − n tanh ∂x ξ First let us change variable from x to the dimensionless variable z = the Bargmann equation takes the form

x ξ.

Then

∂2 2 m Eα ξ 2 2 Φα,n (z) Φ (z) + n ( n + 1 ) sech z Φ (z) = − α,n α,n ∂z 2 ¯2 h (1183) 264

Bargmann Potentials 5 2

V(x) = - n ( n + 1 ) V0 sech x n=1

-5

n=2

-10

n=3

V(x) / V0

0

-15 -2

-1.5

-1

-0.5

0

0.5

1

1.5

2

x

Figure 67: Bargmann potentials, Vn (x) for various values of n. Let us assume that Φα,n−1 (z) satisfies the Bargmann equation of order n − 1. Then on pre-multiplying the eigenvalue equation by   ∂ ˆn = M − n tanh z (1184) ∂z we have  2   ∂ ∂ 2 − n tanh z Φ (z) + n ( n − 1 ) sech z Φ (z) α,n−1 α,n−1 ∂z ∂z 2   2 m Eα ξ 2 ∂ = − − n tanh z Φα,n−1 (z) ∂z ¯h2 (1185) ˆ n depends on n via The Hamiltonian H  2  2 ∂ 2 ˆ n = − ¯h H + n ( n + 1 ) sech z 2 m ξ 2 ∂z 2

(1186)

ˆ n with H ˆ n−1 , we obtain On commuting M ˆn H ˆ n−1 M

ˆ n−1 M ˆn − n = H

¯2 h sech2 z m ξ2

ˆn M ˆn = H



∂ − n tanh z ∂z

 (1187)

265

Thus, we find that ˆn M ˆ n Φα,n−1 = Eα M ˆ n Φα,n−1 H

(1188)

This shows that the wave function Φα,n (z)

ˆ n Φα,n−1 (z) = M   ∂ = − n tanh z Φα,n−1 (z) ∂z

(1189)

ˆ n with the same eigenvalue Eα . is an eigenfunction of H ——————————————————————————————————

4.4.1

Exercise 75

ˆ n satisfies the relation Show that the raising operator M ˆn H ˆ n−1 = H ˆn M ˆn M

(1190)

——————————————————————————————————

4.4.2

Solution 75

In dimensionless units, the raising operator is given by ˆ n = ∂ − n tanh z M ∂z

(1191)

ˆ n containing the n-th order Bargmann potential can be and the Hamiltonian H put into the form   2  ¯h2 ∂ 2 ˆn = − + n ( n + 1 ) sech z (1192) H 2 m ξ2 ∂z 2

ˆn H ˆ n−1 and The identity can be proved by starting with the product M re-writing it as ˆn H ˆ n−1 = H ˆ n−1 M ˆn + [ M ˆn , H ˆ n−1 ] M

(1193)

The commutator can be evaluated as     ¯h2 ∂ ¯2 h ∂2 2 ˆ ˆ [ n [ tanh z , [ Mn , Hn−1 ] = − , n ( n − 1 ) sech z ] + ] 2 m ξ2 ∂z 2 m ξ2 ∂z 2 266

 =

2n(n − 1)

¯h2 2 m ξ2





2

sech z tanh z − n

¯2 h 2 m ξ2



=

= = =

∂2 , tanh z ] ∂z 2

 ¯2 h 2 n ( n − 1 ) sech2 z tanh z 2 m ξ2     ¯h2 ∂ ∂ ¯h2 ∂ ∂ −n [ , tanh z ] − n [ , tanh z ] 2 m ξ 2 ∂z ∂z 2 m ξ2 ∂z ∂z   ¯h2 2n(n − 1) sech2 z tanh z 2 m ξ2     ∂ ¯h2 ∂ ¯h2 2 −n sech z − n sech2 z 2 m ξ 2 ∂z 2 m ξ2 ∂z     ¯h2 ¯h2 ∂ 2 2 n2 sech z tanh z − 2 n sech2 z 2 2 2mξ 2mξ ∂z     2 ¯h ∂ −2n sech2 z − n tanh z 2 m ξ2 ∂z   ¯h2 ˆn −2n sech2 z M (1194) 2 m ξ2 

=

[

Inserting this in the expression (1193), one finds   ¯2 h ˆn H ˆ n−1 = H ˆ n−1 M ˆn − 2 n ˆn M sech2 z M 2 m ξ2 ˆn M ˆn = H

(1195)

which proves the identity. —————————————————————————————————— Starting from n = 0, one can generate all the scattering states by iteration, ˆn M ˆ n−1 . . . M ˆ2 M ˆ 1 Φk,0 (z) Φk,n (z) = M

(1196)

Let us note that at z → + ∞ the solution takes the asymptotic form   m=n  Y  Φk,n (z) = A exp + i k z + ik − m m=1

 + B

− ikz

exp

 m=n Y 

 − ik − m

(1197)

m=1

whereas at z → − ∞ the solution takes the form   m=n  Y  Φk,n (z) = A exp + i k z + ik + m m=1

 + B

exp

− ikz

 m=n Y  m=1

267

 − ik + m

(1198)

On setting B = 0, a case which corresponds to an incident beam with momentum p = h ¯ k, we see that there is no reflected beam and the transmitted beam has a phase which is different from that of the incident beam. The change of phase 2 δ(k) is calculated from δ(k) =

m=n X

−1



tan

m=1

k m

 (1199)

The phase shift δ(k) varies as k is varied. The total variation of the pase shift

Phase shift for the n=4 Bargmann potential 2

δ(k)/π δ( )/π

1.5

n=4 1

0.5

0 0

1

2

3

4

5

6

7

8

ka/π

Figure 68: The k dependence of the phase shift δ(k) for the scattering states of the n = 4 Bargmann potential. is defined as δ(∞) − δ(0)

(1200)

and is equal to n π2 . The total change in the phase shift is related to the number of bound states of the equation through Levinson’s theorem13 . On increasing n by unity, the number of bound states of the Bargmann increases by unity. Thus, the Bargmann equation of order n has n bound states. The lowest energy bound state is given by the un-normalized wave function Φn,n (z) = sechn z 13 N.

Levinson, Kgl. Danske Videnskab. Selskab, Mat.-fys. Medd., 25, 1 (1949).

268

(1201)

Raising Operators for the Bargmann Potentials 2

n=0

1

Ek > 0

0

E =0

n=1 M1

E / V0

E1

n=2

n=3

M2

M3

M2

M3

E2

M3

-1 -2 -3 -4

E3 -5

Figure 69: A graphical description of the raising operators for the Bargmann equations. The raising operators Mn transform the solutions of the (n − 1)-th Bargmann potential to solutions of the n-th potential, but with the same energy eigenvalue. and has a bound state energy given by En = −

¯2 h n2 2 m ξ2

(1202)

Then the higher energy bound states Φm,n (z) can be created from the correh ¯2 2 sponding bound states Φm,m (z) which have energies Em = − 2 m ξ 2 m . The higher-energy bound states are found by operating on Φm,m (z) with the sequenˆn M ˆ n−1 . . . M ˆ n0 . . . M ˆ m+2 M ˆ m+1 (where tially ordered product of operators M 0 n ≥ n ≥ m + 1). Thus, for example, the second lowest energy (un-normalized) eigenfunction of the n-th order Bargmann equation is Φn−1,n (z)

ˆ n Φn−1,n−1 (z) = M   ∂ = − n tanh z sechn−1 z ∂z = − ( 2 n − 1 ) sechn−1 z tanh z

(1203)

which has the energy eigenvalue En−1 = −

¯2 h ( n − 1 )2 2 m ξ2 269

(1204)

——————————————————————————————————

4.4.3

Exercise 76

Find the eigenfunctions and eigenvalues of the Bargmann equation with index n = 2. Also show that the bound states are orthogonal to the scattering states.

The n=2 Bargmann Potential 4

n=2

V(x) / V0

2 0

E1 -2

E0

-4 -6 -8 -3

-2

-1

0

x

1

2

3

Figure 70: The Bargmann potential for n = 2. There are two bound states and the bound state energies are denoted by horizontal lines. ——————————————————————————————————

4.4.4

Solution 76

In terms of the dimensionless variable z, lowest energy bound state of the n-th Bargmann potential has the wave function φn,n (z) = C sechn z

(1205)

and the dimensionless energy eigenvalue is En = − n2 . The higher energy ˆ n acting on the bound bound state is found by using the raising operator M 270

states of the (n − 1)-th Bargmann potential, i.e., ˆ n φm,n−1 (z) φm,n (z) = M

(1206)

The resulting bound state φm,n (z) has energy Em . In this manner, we find that φn−1,n (z)

ˆ n φn−1,n−1 (z) = M   ∂ = − n tanh z sechn−1 z ∂z = − ( 2 n − 1 ) sechn−1 z tanh z

(1207)

The eigenfunction φn−1,n (z) corresponds to the energy eigenvalue En−1 − ( n − 1 )2 . Thus, for n = 2, the two bound states are given by φ2 (z)

= sech2 z

φ1 (z)

= sech z tanh z

=

(1208)

The bound state wave functions are shown in fig(71).

Bound state wave functions for the n = 2 potential 1

n=2

Ψ0(x)

Ψn(x)

0.5

0

Ψ1(x)

-0.5

-1 -4

-3

-2

-1

0

1

2

3

4

x

Figure 71: The bound states of Bargmann potential for n = 2. For n = 2 there are two bound states, the two bound state wave functions are denoted by Ψ0 (x) and Ψ1 (x).

271

The scattering states are given by φk (z) which can be determined from the scattering states of the n = 0 potential, by using the raising operator twice. The scattering states for n = 2 are given by      ˆ ˆ φk (z) = M2 M1 A exp + i k z + B exp − i k z          ∂ ∂ = − 2 tanh z − tanh z A exp + i k z + B exp − i k z ∂z ∂z     2 2 = 2 − k − 3 i k tanh z − 3 sech z A exp + i k z     2 2 + 2 − k + 3 i k tanh z − 3 sech z B exp − i k z (1209) Note that on analytically continuing from k to i κ, one finds that the asymptotic exponentially growing term, for both positive and negative z, has a vanishing coefficient in the asymptotic limit if κ is equal to either 2 or 1. The solutions for these special values of κ correspond to the bound states that we have already found. It remains to show that the bound states are orthogonal to the scattering states. That is, we have to show that Z +∞ dz φm (z) φk (z) = 0 (1210) −∞

For the lowest bound state where m = 2, this requires that coefficients of A and B vanish separately. These both vanish if the integral   Z ∞ Z ∞ dz cos k z 2 − k 2 − 3 sech2 z sech2 z − 3 i k dz sin k z tanh z sech2 z 0

0

(1211) also vanishes. On integrating the second term by parts, we establish that the two terms can be combined as   Z ∞ 1 2 = dz cos k z 2 + k − 3 sech2 z sech2 z (1212) 2 0 As the integrals are evaluated as Z



dz cos k z sech2 z =

0

and Z 0



π k 2 sinh π2k

π k 1 2 dz cos k z sech z = 3! sinh π2k 4

272

 4 + k

(1213)

2

 (1214)

The overlap matrix element vanishes. The orthogonality of the scattering states and the m = 1 bound state is treated similarly. ——————————————————————————————————

4.4.5

Exercise 77

Find the normalization for the two lowest energy bound states of the Bargmann potential for arbitrary n. ——————————————————————————————————

4.4.6

Solution 77

The lowest energy solution of the n-th order Bargmann eigenvalue equation  2   h2 ¯ ∂ n(n + 1) x 2 − + sech φm (x) = Em φm (x) (1215) 2 2 2 m ∂x ξ ξ is seen to be n

φn (x) = A sech



x ξ

 (1216)

This can be most easily seen by changing to the scaled variable z =

x ξ

so the eigenvalue equation becomes  2  ∂ 2 m ξ2 + n ( n + 1 ) φ (z ξ) = − E φn (z ξ) n ∂z 2 ¯2 h

(1217)

(1218)

The function φn (zξ) is an eigenfunction, as can be seen by direct substitution of φn (z ξ) = A sechn z (1219) into the eigenvalue equation and then by noting that ∂2 sechn z ∂z 2

= n ( n + 1 ) sechn z tanh2 z − n sechn z = n2 sechn z − n ( n + 1 ) sech(n+2) z (1220)

Thus, sechn z is a bound state with the bound state energy given by En = −

¯2 h n2 2 m ξ2

273

(1221)

ˆ n on sechn−1 z. Hence, the second bound state φn−1 can be found by applying M   ∂ φn−1 (z) ∝ − n tanh z sechn−1 z (1222) ∂z which leads to the identification of φn−1 = B sechn−1 z tanh z

(1223)

as the bound state with energy eigenvalue En−1 = −

¯2 h ( n − 1 )2 2 m ξ2

(1224)

The magnitude of the normalization constant A can be determined from Z ∞ 2 1 = |A| dx sech2n z −∞ Z ∞ 2 = |A| ξ dz sech2n z (1225) −∞

The integration can be evaluated by integration by parts Z ∞ Z ∞ ∂ dz sech2n dz = dz sech2n−2 z tanh z ∂z −∞ −∞ ∞ Z ∞ ∂ = sech2(n−1) z tanh z − dz tanh z sech2(n−1) z ∂z −∞ −∞ (1226) The boundary term is zero for n ≥ 2 and for n = 1 the boundary term is just 2. Since, ∂ sech2(n−1) z = − 2 ( n − 1 ) sech2(n−1) z tanh z ∂z

(1227)

and tanh2 z = 1 − sech2 z one has the recursion relation Z ∞ dz sech2n dz = −∞

=

2(n − 1) (2n − 1) 2n ( n − 1 )! ( 2 n − 1 )!!

Z

(1228) ∞

dz sech2(n−1) z

−∞

(1229)

for n ≥ 1. Thus, the normalized wave function for the lowest energy bound state is given by s   ( 2 n − 1 )!! x n φn (x) = sech (1230) n ξ 2 ( n − 1 )! ξ 274

The normalization of the next lowest-energy bound state is found from Z ∞ 1 = | B |2 ξ dz sech2(n−1) z tanh2 z −∞  Z ∞  Z ∞ 2(n−1) 2n 2 = |B| ξ dz sech z − dz sech z 

−∞ (n−1)

−∞ n

( n − 2 )! 2 ( n − 1 )! − ( 2 n − 3 )!! ( 2 n − 1 )!!  (n−1)  2 ( n − 2 )! = | B |2 ξ ( 2 n − 1 )!!

= | B |2 ξ

2



Thus, the normalized bound state wave function is given by s     x x ( 2 n − 1 )!! n φn−1 (x) = sech tanh ξ 2n−1 ( n − 2 )! ξ ξ

(1231)

(1232)

——————————————————————————————————

275

4.5

Orbital Angular Momentum

In three dimensions, the orbital angular momentum operator is a pseudo-vector and is defined by the vector product ˆ = rˆ ∧ pˆ L

(1233)

therefore, it can be decomposed in terms of Cartesian unit vectors via ˆ = eˆx L ˆ x + eˆy L ˆ y + eˆz L ˆz L

(1234)

where the components are given by the expressions ˆx L ˆy L

= yˆ pˆz − zˆ pˆy

ˆz L

= x ˆ pˆy − yˆ pˆx

= zˆ pˆx − x ˆ pˆz (1235)

From the commutation relations between position and momentum operators, one finds that the components of the angular momentum operator satisfy the following type of commutation relations with the components of the position and momentum ˆ x , yˆ ] = i ¯h zˆ [L ˆ x , pˆy ] = i ¯h pˆz [L ˆx , x [L ˆ] = 0 ˆ x , pˆx ] = 0 [L

(1236)

The above expressions can be used together with the rules for the commutators of sums and products of operators, to prove that ˆx , [L ˆy , [L

ˆ y ] = i ¯h L ˆz L ˆz ] = i h ˆx L ¯ L ˆz , L ˆ x ] = i ¯h L ˆy [L

(1237)

Since the components of the orbital angular momentum operators do not commute, the uncertainty principle asserts that, in general, it is not possible to find an a simultaneous eigenstate of more than one component. Since the angular momentum operators satisfy the above commutation relations, they provide an example of a Lie algebra14 . 14 The

operators of a Lie algebra satisfy commutation relations of the form ˆi , A ˆj ] = [A

X

k ˆk Ci,j A

k

ˆi which form a Lie algebra are linearly related to the infinitesimal The set of operators A generators of transformation groups, known as Lie Groups. In the case of orbital angular momentum, the Lie group is the group of three-dimensional rotations. It is also noteworthy

276

The commutation relations between the various components of the angular momentum can be evaluated through ˆx , L ˆy ] [L

ˆ x , z pˆx ] − [ L ˆ x , x pˆz ] = [L ˆ x , z ] pˆx + z [ L ˆ x , pˆx ] − [ L ˆ x , x ] pˆz − x [ L ˆ x , pˆz ] = [L = −i¯ h y pˆx + i ¯h x pˆy ˆz = i¯ hL (1238)

The others can be obtained by cyclic permutations, corresponding to the different choices of Cartesian coordinate axes. The operator expressing the squared magnitude of the angular momentum vector follows from the operators representing the Cartesian components and the Pythagorean theorem c2 = L ˆ 2x + L ˆ 2y + L ˆ 2z L

(1239)

That is, the squared magnitude of the angular momentum is merely the sum of the squares of the operator components. The magnitude of L2 is a scalar quantity. It can be proved that c2 ] = 0 ˆx , L [L c2 ] = 0 ˆy , L [L c2 ] = 0 ˆ , L [L z

(1240)

Thus, it is possible to find simultaneous eigenfunctions of any one component of the angular momentum and the magnitude squared. The commutation relation between the z component of the angular momenˆ z and the magnitude of the angular momentum L ˆ 2 is proved by considering tum L ˆz , L b2 ] = [ L ˆz , L ˆ 2x ] + [ L ˆz , L ˆ 2y ] + [ L ˆz , L ˆ 2z ] [L

(1241)

that the set of commutation relations between the angular momentum operators and any vector quantity with components vˆi satisfy commutation relations of the form ˆ i , vˆj ] = i ¯ [L h

X

i,j,k vˆk

k

where i,j,k is the Levi-Civita symbol. The Levi-Civita symbol is defined as

 i,j,k =

1 −1 0

if i, j, k is a cyclic permutation of 1, 2, 3 if i, j, k is not a cyclic permutation of 1, 2, 3 otherwise

The above commutation relations imply that all vectors transform in the same way under the set of rotation operations.

277

which becomes ˆz , L b2 ] [L

ˆz , L ˆ 2x ] + [ L ˆz , L ˆ 2y ] = [L ˆz , L ˆx ] L ˆx + [ L ˆz , L ˆx ] L ˆx + [ L ˆz , L ˆy ] L ˆy + L ˆy [ L ˆz , L ˆy ] = [L     ˆy L ˆx + L ˆx L ˆy ˆx L ˆy + L ˆy L ˆx = i¯ h L − i ¯h L =

0

(1242)

ˆ z and L b 2 commute. By invariance under the permutation of the coorThus, L b 2 with the other components of L ˆ are also dinate axes, the commutators of L found to be zero. ——————————————————————————————————

4.5.1

Exercise 78

Find the expression for the operator L2 in terms of the Cartesian components of the position and derivatives w.r.t position. Express your final result entirely in terms of invariant quantities, such as scalar products. ——————————————————————————————————

4.5.2

Solution 78

The components of the angular momentum can be written in terms of the LeviCivita symbol i,j,k as Lk = − i ¯h

X

i,j,k xi

i,j

∂ ∂xj

(1243)

where i,j,k = 1

(1244)

if i, j, k are an even permutation of the indices 1, 2, and 3, and i,j,k = − 1

(1245)

if i, j, k are an odd permutation of the indices 1, 2, and 3, and i,j,k = 0 if any index is repeated.

278

(1246)

The magnitude of the angular momentum is given by c2 = X L ˆ2 L k

k

= −h ¯2

X X

i,j,k xi

i,j

k

∂ ∂ X l,m,k xl ∂xj ∂xm l,m

(1247) But on using the identity X i,j,k l,m,k = δi,l δj,m − δi,m δj,l

(1248)

k

one obtains c2 L

= − ¯h

2

X 

∂ ∂ ∂ ∂ xi − xi xj xi ∂xj ∂xj ∂xj ∂xi

i,j

= −h ¯2

X 

xi xi

i,j

=



∂ ∂xi ∂ ∂ + xi ∂xj ∂xj ∂xj ∂xj

∂ ∂ ∂xj ∂ − xi − xi xj ∂xj ∂xi ∂xj ∂xi 2 X  ∂ ∂ −¯ h2 x2i + xi δi,j ∂x2j ∂xj i,j  ∂ ∂ ∂ − xi xj − xi ∂xj ∂xi ∂xi



(1249)

Thus, we have c2 L

= −h ¯2

 X

x2i

i



X i,j

X ∂2 X ∂ + xi 2 ∂x ∂x i j j i

 X ∂ ∂ ∂ xi xj − 3 xi ∂xj ∂xi ∂xi i

(1250)

Hence, the magnitude of the angular momentum can be written as  X X ∂2 X ∂ c 2 2 L = −¯ h x2i + xi 2 ∂xj ∂xi i j i  X X ∂ ∂ ∂ − xi xj − 3 xi ∂xj ∂xi ∂xi i,j i  X  X ∂2 X X ∂ ∂ ∂ = −¯ h2 x2i − x x − 2 x i i j ∂x2j ∂xj ∂xi ∂xi i j i,j i (1251) 279

which can be expressed as  X  X ∂2 X X ∂ ∂ ∂ c 2 2 2 L = −¯ h xi − xj xi − 1 xi ∂x2j ∂xj ∂xi ∂xi i j i,j i (1252) When put in vector notation, the magnitude of the angular momentum becomes   c2 = − ¯ L h2 r2 ∇2 − ( r . ∇ )2 − ( r . ∇ ) (1253) which is seen to be a scalar operator. ——————————————————————————————————

4.5.3

Exercise 79

c2 and find the ˆ z and L Show that the following functions are eigenfunctions of L corresponding eigenvalues. Ψ0 (r) Ψ1 (r) Ψ2 (r) Ψ3 (r) Ψ4 (r) Ψ5 (r) Ψ6 (r) Ψ7 (r) Ψ8 (r)

= = = = = = = = =

g0 (r) z g1 (r) (x + i (x − i ( 3 z2 − (x + i (x − i z(x + z(x −

y ) g2 (r) y ) g3 (r) r2 ) g4 (r) y )2 g5 (r) y )2 g6 (r) i y ) g7 (r) i y ) g8 (r)

(1254)

where the functions gn (r) are arbitrary functions of the radial distance. —————————————————————————————————— The Cartesian components of the angular momentum operators can be expressed in terms of spherical polar coordinates. The Cartesian components are found to be given by the expressions   ∂ ∂ ˆ Lx = − i ¯ h − sin ϕ − cos ϕ cot θ ∂θ ∂ϕ   ∂ ∂ ˆ Ly = − i ¯h + cos ϕ − sin ϕ cot θ ∂θ ∂ϕ ∂ ˆ z = − i ¯h L (1255) ∂ϕ 280

and these components can be shown to satisfy the same commutation relations that were derived from a purely Cartesian formulation. We shall commence the derivation of eqns(1255) by expressing the angular momentum in terms of the unit vectors of the polar coordinate system. The orbital angular momentum operator is expressed as ˆ = r ∧ pˆ L = − i ¯h r ∧ ∇

(1256)

and on using the representation ∇ = eˆr

∂ 1 ∂ 1 ∂ + eˆθ + eˆϕ ∂r r ∂θ r sin θ ∂ϕ

(1257)

one finds   1 ∂ 1 ∂ ˆ L = −i¯ h r eˆr ∧ eˆθ + eˆr ∧ eˆϕ r ∂θ r sin θ ∂ϕ   1 ∂ ∂ − eˆθ = − i ¯h eˆϕ ∂θ sin θ ∂ϕ

(1258)

In the last line we have used the fact that the unit vectors eˆr , eˆθ and eˆϕ form an orthogonal coordinate system. Furthermore, as ˆr = eˆr r  = r cos θ eˆz + r sin θ

 sin ϕ eˆy + cos ϕ eˆx

(1259)

one can express the spherical polar coordinate unit vectors in terms of the Cartesian unit vectors by using the definition of the unit vectors eˆr

=

∂ r ∂r

=

cos θ eˆz + sin θ

=

1 ∂ r r ∂θ



eˆθ

 sin ϕ eˆy + cos ϕ eˆx

 = − sin θ eˆz + cos θ eˆϕ

 sin ϕ eˆy + cos ϕ eˆx

1 ∂ r r sin θ ∂ϕ   = cos ϕ eˆy − sin ϕ eˆx =

(1260)

On substituting the above expressions for the unit vectors into the equation for ˆ in eqn(1258), the angular momentum is found the angular momentum vector L 281

in the form ˆ = −i¯ L h

 − eˆx sin ϕ

∂ ∂ + eˆy cos ϕ ∂θ ∂θ

∂ ∂ ∂ − eˆx cot θ cos ϕ − eˆy cot θ sin ϕ + eˆz ∂ϕ ∂ϕ ∂ϕ

 (1261)

which has three components given by the expressions of eqn(1255). c2 is given by The square of the angular momentum L    1 ∂2 1 ∂ ∂ c 2 2 L = − ¯h + sin θ sin θ ∂θ ∂θ sin2 θ ∂ϕ2

4.5.4

(1262)

Simultaneous Eigenfunctions.

In spherical polar coordinates, the angular momentum operators act on the wave functions Ψ(r, θ, ϕ), but since r does not appear in the operators, it is c2 ˆ z and L redundant. The simultaneous eigenfunctions of the pair of operators L l are functions of θ and ϕ alone, and are written as Ym (θ, ϕ). These eigenfunctions satisfy the pair of eigenvalue equations,    ∂2 1 ∂ ∂ 1 c 2 2 l + sin θ Yml (θ, ϕ) L Ym (θ, ϕ) = − ¯ h sin θ ∂θ ∂θ sin2 θ ∂ϕ2

ˆ z Y l (θ, ϕ) L m

= λl Yml (θ, ϕ)

(1263)

∂ Y l (θ, ϕ) ∂ϕ m = µ Yml (θ, ϕ)

(1264)

= −i¯ h

c2 and µ is the eigenvalue of L ˆ z . The eigenfunctions where λl is the eigenvalue of L l Ym (θ, ϕ) are usually factorized according to Yml (θ, ϕ) = Θlm (θ) Φm (ϕ)

(1265)

ˆ z operator as a derivative w.r.t. ϕ, one can easily find From the form of the L that the ϕ dependence of Yml (θ, ϕ) must be given by Yml (θ, ϕ) = Θlm (θ)



1 2π

 12

 exp

i

µϕ ¯h

 (1266)

The modulus of the wave function is single valued, as it must represent a unique probability density for each point in space. Therefore, we expect that the wave

282

function at the point (r,θ,ϕ) must have the same values at (r,θ,ϕ + 2π) since this represents the same point. In this case, we have     µ(ϕ + 2π) µϕ exp i = exp i (1267) ¯h ¯h which is satisfied if µ = h ¯ m for any positive or negative integer m. Classically, the magnitude of a component of the angular momentum should be smaller than the magnitude of the pseudo-vector L. From this observation, one expects that an inequality should exist between the eigenvalues of λl and m¯ h. The maximum value of m will be denoted by l. Later, we shall show that the eigenvalue λl is related to l via λl = h ¯ 2 l ( l + 1 ). This suggests a picture of the orbital angular √ momentum eigenstates as beˆ has a definite magnitude λl and a definite z component ing states where L m¯ h. The picture is that in which the vector L has a definite z component and magnitude, but has an uncertain direction due to precession around the z axis.

m

3

2

1

[ l ( l + 1 ) ]1/2 0

-1

-2

-3

Figure 72: A semi-classical picture of a state with angular momentum l and z component m. The picture of the angular momentum is a vector of length p ¯h l ( l + 1 ) which has a projection h ¯ m along the z axis. The angular momentum vector can be thought of as precessing around the z axis, so that the x and y components are indeterminate.

283

4.5.5

The Raising and Lowering Operators

ˆ + and Two useful operators are given by the raising and lowering operators, L ˆ L− defined by ˆ+ L ˆ− L

ˆx + i = L ˆx − i = L

ˆy L ˆy L

(1268)

ˆ + and L ˆ − are Hermitean conjugates. These operators also The operators L c 2 commute with L . c2 , L c2 , L ˆ+ ] = [ L ˆ− ] = 0 [L

(1269)

c2 . ˆ x and L ˆ y commute with L as both L In spherical polar coordinates, the raising and lowering operators are given by ˆ+ L ˆ− L



 ∂ ∂ + i cot θ ∂θ ∂ϕ    ∂ ∂ = −¯ h exp − i ϕ − i cot θ ∂θ ∂ϕ

=

+h ¯ exp



+ iϕ

(1270)

The raising and lowering operators satisfy the commutation relations ˆz , L ˆ+ ] = + h ˆ+ [L ¯ L ˆz , L ˆ − ] = − ¯h L ˆ− [L ˆ ˆ ˆ [ L+ , L− ] = 2 h ¯ Lz

(1271)

c2 , L c2 , ˆ ± ] on the eigenfunctions of L On operating with the commutators [ L one finds c2 L c2 Y l (θ, ϕ) ˆ ± Yml (θ, ϕ) = L ˆ± L L m ˆ ± Yml (θ, ϕ) = λl L

(1272)

c2 with the same ˆ ± Y l (θ, ϕ) is also an eigenfunction of L and recognizes that L m l ˆ ± acting on an eigenfunction of the eigenvalue λl as found for Ym (θ, ϕ). Thus, L magnitude of the orbital angular momentum does not change the eigenvalue λl . ˆz , L ˆ± On operating with the commutator [ L l functions Ym (θ, ϕ), one finds  l ˆ ˆ ˆ± L ˆ z ± ¯h Lz L± Ym (θ, ϕ) = L   = m ¯h ± ¯h

284

ˆ ± on the eigen] = ± ¯h L

ˆ± L



Yml (θ, ϕ)

ˆ ± Yml (θ, ϕ) L

(1273)

Thus, the raising and lowering operators when acting on a simultaneous eigenc2 and L ˆ z produce other simultaneous eigenfunctions with L ˆ z eigenfunction of L ˆ values that are either raised or lowered by h ¯ . The action of L± on the eigenl function Yml (θ, ϕ) is to produce a function proportional to Ym±1 (θ, ϕ). Thus, we have l ˆ ± Yml (θ, ϕ) = C± (l, m) Ym±1 L (θ, ϕ) (1274) The constants of proportionality, C± (l, m), have yet to be determined.

4.5.6

The Eigenvalues and Degeneracy

First we shall note the two equalities, ˆ− L ˆ+ L ˆ+ L ˆ− L

c2 − L ˆ 2z − ¯h L ˆz = L c2 − L ˆ 2z + ¯h L ˆz = L

(1275)

ˆ− L ˆ+ We shall first consider the effect of the raising operator. On operating L l on the eigenfunction Ym (θ, ϕ), on using the first equality of eqn(1275), one finds   c2 − L ˆ− L ˆ + Yml (θ, ϕ) = ˆ 2z − ¯h L ˆ z Yml (θ, ϕ) L L   = λl − ¯h2 m2 − ¯h2 m Yml (θ, ϕ) (1276) If the value of m is the maximum value l, then we have ˆ + Yll (θ, ϕ) = 0 L

(1277)

and so, on substituting m = l in the previous equation, we find   2 λl = h ¯ l l + 1

(1278)

Thus, we have found the eigenvalue of the square of the angular momentum in terms of the maximum eigenvalue of the z component of the angular momentum operator. Now we shall determine explicit expressions for the products of the coefficients C± (l, m). On using the properties of the raising and lowering operators given by eqn(1274), one finds ˆ− L ˆ + Y l (θ, ϕ) = C− (l, m + 1) C+ (l, m) Y l (θ, ϕ) L m m

(1279)

but we also have ˆ− L ˆ + Yml (θ, ϕ) = L



2

2

λl − ¯h m 285

2

− ¯h m



Yml (θ, ϕ)

(1280)

Therefore, we can determine the products of the coefficient C± (l, m) by equating the above two expressions and then by substituting the expression for λl . Thus, we find   C− (l, m + 1) C+ (l, m) = λl − ¯h2 m2 + m   = h ¯ 2 l ( l + 1 ) − m2 − m (1281)

The effect of the lowering operator is found by considering the other equality, ˆ+ L ˆ − on the eigenfuncand considering the minimum value of l. On operating L tion Yml (θ, ϕ) and on using the second equality of eqn(1275), one finds   c 2 2 l ˆ ˆ ˆ ˆ L+ L− Ym (θ, ϕ) = L − Lz + ¯h Lz Yml (θ, ϕ)   2 2 = h ¯ l ( l + 1 ) − m + m Yml (θ, ϕ) (1282) If m is the minimum value l0 , then we have ˆ − Yll0 (θ, ϕ) = 0 L therefore ¯2 h



l ( l + 1 ) − l02 + l0

(1283)  = 0

(1284)

so we find that the minimum value of m is given by l0 = − l. Thus, the possible values of m run through the set of numbers − l , − ( l − 1 ) , − ( l − 2 ) , . . . , ( l − 2 ) , ( l − 1 ) , l , which includes 0. There are ( 2 l + 1 ) c2 corresponddifferent eigenfunctions of m for fixed l. Thus, the eigenvalues of L ing to the value l have a degeneracy of ( 2 l + 1 ). Since we have proved that the eigenvalues m for the z component of the orbital angular momentum are integer, we note that l must also be an integer.

4.5.7

The Effect of the Raising Operators.

ˆ + and L ˆ − are Hermitean The coefficients C± (l, m) are found by noting that as L conjugates Z π Z 2π l ˆ + Yml (θ, ϕ) dθ sin θ dϕ Ym+1 (θ, ϕ)∗ L 0

0

= C+ (l, m) Z π Z = dθ sin θ 0

2 π

ˆ − Y l (θ, ϕ) dϕ Yml (θ, ϕ)∗ L m+1

∗

0

∗ = C− (l, m + 1)

(1285) 286

Thus, we have | C+ (l, m) |2

= h ¯2



 l(l + 1) − m(m + 1)

= h ¯2 ( l − m ) ( l + m + 1 )

(1286)

and, by considering the lowering operator, one finds the analogous equation   | C− (l, m) |2 = h ¯2 l ( l + 1 ) − m ( m − 1 ) = h ¯2 ( l + m ) ( l − m + 1 )

(1287)

Thus, the effect of the raising and lowering operators are given by p l ˆ + Yml (θ, ϕ) = h L ¯ ( l − m ) ( l + m + 1 ) Ym+1 (θ, ϕ) p l l ˆ L− Ym (θ, ϕ) = h ¯ ( l + m ) ( l − m + 1 ) Ym−1 (θ, ϕ) (1288) where we have chosen the phase of the constants to be zero.

4.5.8

Explicit Expressions for the Eigenfunctions

Explicit expressions for the simultaneous eigenfunctions Yml (θ, ϕ) can be found starting from the effect of the raising operator on the state with maximum m, i.e. m = l. This results in the equation ˆ + Yll (θ, ϕ) = 0 L which has the explicit form    ∂ ∂ + i cot θ Yll (θ, ϕ) = 0 h exp + i ϕ ¯ ∂θ ∂ϕ

(1289)

(1290)

As the eigenfunction has a ϕ dependence of exp[ i l ϕ ], and as the eigenfunction is factorized as    1 1 2 l l exp i m ϕ (1291) Ym (θ, ϕ) = Θm (θ) 2π one finds

∂ Θl (θ) = l cot θ Θll (θ) ∂θ l This equation can be integrated to yield Θll (θ) = Bll sinl θ

(1292)

(1293)

where Bll is a constant of proportionality. The magnitude of Bll is determined by the normalization condition Z π l 2 | Bl | dθ sin θ sin2l θ = 1 (1294) 0

287

Thus, the magnitude of the normalization is found as |

Bll

 | =

( 2 l + 1 )! 22l+1 ( l! )2

 12 (1295)

The normalized eigenfunction Yll (θ, ϕ) is found as Yll (θ, ϕ)

 =

1 ( 2 l + 1 )! 4 π 22l ( l! )2

 12



l

sin θ exp

 ilϕ

(1296)

up to an arbitrary constant phase. The lowering operator acting on the state with minimum m, i.e. m = − l, results in the equation l ˆ − Y−l L (θ, ϕ) = 0

(1297)

which has the explicit form    ∂ ∂ h exp − i ϕ ¯ − + i cot θ Yll (θ, ϕ) = 0 ∂θ ∂ϕ

(1298)

As the eigenfunction has a ϕ dependence of exp[ − i l ϕ ], and as the eigenfunction is factorized as  1   1 2 l l Ym (θ, ϕ) = Θm (θ) exp i m ϕ (1299) 2π one finds

∂ Θl (θ) = l cot θ Θl−l (θ) ∂θ −l which can be integrated to yield

(1300)

l Θl−l (θ) = B−l sinl θ

(1301)

l l is determined where B−l is a constant of proportionality. The magnitude of B−l by the normalization condition Z π l | B−l |2 dθ sin θ sin2l θ = 1 (1302) 0

Thus, the magnitude of the normalization is found as |

l B−l

 | =

( 2 l + 1 )! 22l+1 ( l! )2

 12 (1303)

l The normalized eigenfunction Y−l (θ, ϕ) is found as

l Y−l (θ, ϕ)

 =

1 ( 2 l + 1 )! 4 π 22l ( l! )2

 12

288

l

sin θ exp



 − ilϕ

(1304)

up to an arbitrary constant phase. The eigenfunctions corresponding to larger values of m can be found by successive operation with the raising operators,    ∂ ∂ exp + i ϕ + i cot θ Yml (θ, ϕ) ∂θ ∂ϕ p l = ( l − m ) ( l + m + 1 ) Ym+1 (θ, ϕ) (1305) Thus, on factorizing the eigenfunctions, we have   p ∂ − m cot θ Θlm (θ) = ( l − m ) ( l + m + 1 ) Θlm+1 (θ) + ∂θ   p ∂ 1 sinm θ Θlm (θ) = ( l − m ) ( l + m + 1 ) Θlm+1 (θ) m ∂θ sin θ   p ∂ 1 sinm+1 θ Θlm (θ) = ( l − m ) ( l + m + 1 ) Θlm+1 (θ) m ∂ cos θ sin θ (1306) Starting with m = − l, after l + m iterations, one finds l+m  1  ( − 1 )l+m ( 2 l + 1 ) ( l − m )! 2 ∂ m l sin2l θ Θm (θ) = sin θ ∂ cos θ 2l l! 2 ( l + m )! (1307) Thus, the spherical harmonics are given by  1 ( − 1 )l+m ( 2 l + 1 ) ( l − m )! 2 Yml (θ) = × 2l l! 4 π ( l + m )!  l+m   ∂ m 2l × sin θ sin θ exp i m ϕ (1308) ∂ cos θ These eigenfunctions are normalized via Z 2π Z π dϕ dθ sin θ | Yml (θ, ϕ) |2 = 1 0

(1309)

0

The eigenfunction for positive m and negative m are related via l Y−m (θ, ϕ) = ( − 1 )m Yml (θ, ϕ)∗

(1310)

c2 . This can The spherical harmonics also satisfy eigenvalue equation for L be seen explicitly from the representation in spherical polar coordinates    1 ∂2 1 ∂ ∂ −¯ h2 + sin θ Yml (θ, ϕ) = h ¯ 2 l ( l + 1 ) Yml (θ, ϕ) sin θ ∂θ ∂θ sin2 θ ∂ϕ2 (1311) 289

which on factorizing the eigenfunctions into polar and azimuthal parts, yields    m2 1 ∂ ∂ + − sin θ Θlm (θ) = l ( l + 1 ) Θlm (θ) (1312) sin θ ∂θ ∂θ sin2 θ l which identifies Θlm (θ) with the associated Legendre Polynomials Pm (cos θ). This identification proceeds through changing variables x = cos θ, so one finds   ∂ ∂Θ m2 ( 1 − x2 ) − Θ = −l(l + 1)Θ (1313) ∂x ∂x 1 − x2

which can then be put into the standard form ( 1 − x2 )

∂2Θ ∂Θ m2 − 2 x − Θ = −l(l + 1)Θ ∂x2 ∂x 1 − x2

(1314)

The spherical harmonics are given in terms of the associated Legendre Polynomials via s   2 l + 1 ( l − m )! l m l ( − 1 ) Pm ( cos θ ) exp i m ϕ Ym (θ, ϕ) = 4π ( l + m )! (1315)

4.5.9

Legendre Polynomials

The generating function F (z, t) for the Legendre polynomials has an expansion given by F (z, t) = √

1 1 + t2 − 2 z t ∞ X = tl Pl (z)

(1316)

l = 0

for t < 1. The Legendre polynomials satisfy recurrence relations, which can be derived from the generating function expansion. The first is obtained by differentiating the generating function with respect to t   l=∞ X ∂F (z, t) (z − t) = = l tl−1 Pl (z) (1317) 3 2 ∂t ( 1 + t − 2 z t )2 l=0 After multiplying the above equation with a factor of ( 1 + t2 − 2 z t ) and then substituting the generating function expansion, one obtains the relation ( 1 + t2 − 2 z t )

l=∞ X

l tl−1 Pl (z) = ( z − t )

l=0

l=∞ X l=0

290

tl Pl (z)

(1318)

z 1

|Ylm(θ,φ)|2 0.5

θ

0 -1

-0.5

0

0.5

1

-0.5

-1

Figure 73: Since the spherical harmonics Yml (θ, ϕ) only depend on ϕ through a phase factor, their θ dependence can be conveniently represented by a polar plot. In a polar  plot, pointson a curve are represented by the radial and angular coordinates

|Yml (θ, ϕ)|2 , θ .

On equating like powers of t, one has ( l + 1 ) Pl+1 (z) − ( 2 l + 1 ) z Pl (z) + l Pl−1 (z) = 0

(1319)

A second recurrence relation can be found by differentiating the generating function expansion with respect to z 

∂F (z, t) ∂z

 =

t 3

( 1 + t2 − 2 z t ) 2

=

l=∞ X

tl Pl0 (z)

(1320)

l=0

On multiplying by ( 1 + t2 − 2 z t ) and then substituting the generating function expansion, one obtains X X t tl Pl (z) = ( 1 + t2 − 2 z t ) tl Pl0 (z) (1321) l

l

which yields the relation 0 0 Pl+1 (z) + Pl−1 (z) = 2 z Pl0 (z) + Pl (z)

291

(1322)

involving the derivatives of the polynomials. Numerous other relations can be obtained from the above recurrence relations. If we take the derivative of the first recurrence relation, eqn(1319), then we have 0 0 ( l + 1 ) Pl+1 (z) − ( 2 l + 1 ) z Pl0 (z) + l Pl−1 (z) = ( 2 l + 1 ) Pl (z) (1323)

The term proportional to z Pl0 (z) can be eliminated from the above equation. This is achieved by multiplying eqn(1323) by two and then subtracting ( 2 l + 1 ) times eqn(1322), which leads to 0 0 Pl−1 (z) − Pl+1 (z) = − ( 2 l + 1 ) Pl (z)

(1324)

A more useful set of equations can be obtained by relating the Legendre polynomials of one order to the next. For example, adding eqn(1322) and eqn(1324) 0 one eliminates Pl+1 (z) 0 Pl−1 (z) = z Pl0 (z) − l Pl (z)

(1325)

On subtracting the same pair of the equations, one has 0 Pl+1 (z) = ( l + 1 ) Pl (z) + z Pl0 (z)

(1326)

Another pair of equations can be obtained from eqn(1325) and eqn(1326) by shifting the index of eqn(1325) from l − 1 to l and then eliminating the term 0 proportional to Pl+1 (z), leading to the equation ( 1 − z 2 ) Pl0 (z) = ( l + 1 )



 z Pl (z) − Pl+1 (z)

(1327)

Another recurrence relation can be derived from the above equation by using eqn(1319) to eliminate the term proportional to ( l + 1 ) Pl+1 (z). This process results in ( 1 − z 2 ) Pl0 (z) = l Pl−1 (z) − l z Pl (z) (1328)

From these relations we can find the differential equation for the Legendre polynomials Pl (z). By differentiating eqn(1328), one obtains an equation in0 volving Pl−1 (z) which can be eliminated using eqn(1325). This procedure leads to Legendre’s equation ( 1 − z 2 ) Pl00 (z) − 2 z Pl0 (z) = − l ( l + 1 ) Pl (z) 292

(1329)

On setting z = cos θ, the differential equation takes the form   1 ∂ ∂ sin θ Pl ( cos θ ) + l ( l + 1 ) Pl ( cos θ ) = 0 sin θ ∂θ ∂θ

(1330)

This is recognized as the eigenvalue equation for the total angular momentum, in the case where there is a simultaneous zero eigenvalue for the z component of the angular momentum, m = 0. In the general case, where m 6= 0, the simultaneous eigenfunctions were denoted by Θlm (θ). For finite m, the θ dependence of the angular momentum eigenvalues is expressed in terms of the associated Legendre polynomials.

4.5.10

Associated Legendre Functions

l The Associated Legendre functions, Pm (z) for positive m, are defined as m

l Pm (z) = ( 1 − z 2 ) 2

∂m Pl (z) ∂z m

(1331)

Clearly, the associated Legendre function with m = 0 is identical with the Legendre polynomial. Furthermore, since Pl (z) is a polynomial of order l, the associated Legendre functions vanish for m > l. The associated Legendre functions satisfy a differential equation obtained by differentiating Legendre’s equation m times. This leads to ( 1 − z 2 ) v 00 − 2 z ( m + 1 ) v 0 + ( l − m ) ( l + m + 1 ) v = 0 (1332) where v =

∂m Pl (z) ∂z m

(1333)

l The associated Legendre function Pm (z) and v are related via

v =

1 l (z) m P ( 1 − z2 ) 2 m

(1334)

The first and second derivatives of v are evaluated as v0 = m

∂ z 1 P l (z) v + m 1 − z2 ( 1 − z 2 ) 2 ∂z m

(1335)

and v 00

 = m +

 z2 1 v + v ( 1 − z 2 )2 1 − z2 1 ∂ 1 ∂2 l Pm (z) + P l (z) m m 2 2 ( 1 − z ) 2 ∂z ( 1 − z ) 2 ∂z 2 m (1336)

(m + 2) 2mz 1 − z2

293

Substitution of the expressions for v, v 0 and v 00 into eqn(1332) leads to   ∂2 l ∂ l m2 l ( 1 − z 2 ) 2 Pm (z) − 2 z Pm (z) + l ( l + 1 ) − Pm (z) = 0 ∂z ∂z 1 − z2 (1337) which is the differential equation satisfied by the associated Legendre functions. On changing variables from z to cos θ, one finds that the associated Legendre polynomials satisfy     1 ∂ ∂ m2 l l sin θ P ( cos θ ) + l ( l + 1 ) − Pm ( cos θ ) = 0 sin θ ∂θ ∂θ m sin2 θ (1338) The above equation occurs as part of eigenvalue equation for the magnitude of the orbital angular momentum, when the ϕ dependence has been separated out by the introduction of the quantum number m corresponding to the z component of the orbital angular momentum. The normalized part of the wave function which depends on the polar angle θ is given by s ( 2 l + 1 ) ( l − m )! l l Θm (θ) = P (cos θ) (1339) 2 ( l + m )! m which involves the associated Legendre function in the variable cos θ.

4.5.11

Spherical Harmonics

Explicit expressions for the first few spherical harmonics Yml (θ, ϕ) are given in Tables(3-4). The spherical harmonics are normalized so that Z

2 π

Z dϕ

0

π

0

dθ Yml 0 (θ, ϕ)∗ Yml (θ, ϕ) = δl,l0 δm,m0

0

294

(1340)

Table 3: The Lowest Order Spherical Harmonics Yml (θ, ϕ). Yml (θ, ϕ)

l

m

l=0

m=0

Y00 (θ, ϕ)

l=1

m=0

Y01 (θ, ϕ)

m = ±1

1 Y±1 (θ, ϕ)

m=0

Y02 (θ, ϕ)

q

m = ±1

2 Y±1 (θ, ϕ)



m = ±2

2 Y±2 (θ, ϕ)

l=2

√1 4 π

q



q

5 16 π

3 4 π

cos θ



3 8 π



i ϕ

sin θ



2

3 cos θ − 1

q

15 8 π

q

15 32 π

e± e±

i ϕ

cos θ sin θ

i 2 ϕ

sin2 θ

Table 4: The Spherical Harmonics with l = 3, Ym3 (θ, ϕ). l

l=3

Yml (θ, ϕ)

m

m=0

Y03 (θ, ϕ)

m = ±1

3 Y±1 (θ, ϕ)

m = ±2

3 Y±2 (θ, ϕ)

m = ±3

3 Y±3 (θ, ϕ)

q



q

7 16 π

5 cos θ − 3 cos θ i ϕ

( 5 cos2 θ − 1 ) sin θ

105 32 π



i 2 ϕ

q

35 64 π





295



3



21 64 π

q



cos θ sin2 θ

i 3 ϕ

sin3 θ

Figure 74: The angular dependence of the squared modulus of the spherical harmonic | Y00 (θ, ϕ) |2 . Since the spherical harmonic does not depend on (θ, ϕ) via the exponential term exp[ i m ϕ ], the angular distribution is spherically symmetric. The θ dependence for l = 0 is shown as a polar plot.

l=0

0

l Θ 0(θ θ) l

m=0

296

2

Figure 75: The angular dependence of the squared modulus of the spherical harmonics | Yml (θ, ϕ) |2 . Since the spherical harmonics only depend on ϕ via the exponential term exp[ i m ϕ ], the probability density is independent of ϕ. The angular dependence is solely determined by the associated Legendre functions Θlm (θ). The θ dependence for l = 1, m = ± 1 , 0 are shown as polar plots.

l=1

1

l Θ 0(θ θ) l

m=0

l=1

1

l Θ 1(θ θ) l

m=1

297

2

2

Under inversion r → − r, then as ϕ → ϕ + π and θ → π − θ, one finds that the spherical harmonics transform as Yml (θ, ϕ) → Yml (π − θ, π + ϕ) → ( − 1 )l Yml (θ, ϕ)

(1341)

Thus, all orbital angular momentum wave functions with the same value of l must have the same parity. The spherical harmonics are simultaneous eigenstates of two Hermitean operators, therefore, they form a complete set. The completeness relation in the space of (θ,ϕ) can be expressed as ∞ l X X

Yml (θ0 , ϕ0 )∗ Yml (θ, ϕ) = δ( rˆ0 − rˆ )

(1342)

l=0 m=−l

where the delta function expresses the condition that the directions rˆ and rˆ0 must be the same. ——————————————————————————————————

4.5.12

Exercise 80

Calculate the vector probability current density j(r, θ, ϕ) for a wave function of the form Ψ(r, θ, ϕ) = f (r) Yml (θ, ϕ) (1343) which has an orbital angular momentum of l = 1, for the various eigenvalues of the z component of the angular momentum, m, if f (r) is a real function. ——————————————————————————————————

4.5.13

Solution 80

The velocity is expressed as v

=

pˆ me

= −i

¯h ∇ me

(1344)

The gradient is written as ∇ = eˆr

∂ 1 ∂ 1 ∂ + eˆϕ + eˆθ ∂r r ∂θ r sin θ ∂ϕ

298

(1345)

l=2 m=0

2

l Θ 0(θ θ) l

2

Figure 76: The angular dependence of the squared modulus of the spherical harmonics | Yml (θ, ϕ) |2 . Since the spherical harmonics only depend on ϕ via the exponential term exp[ i m ϕ ], the probability density is independent of ϕ. The angular dependence is solely determined by the associated Legendre functions Θlm (θ). The θ dependence for l = 2 , m = ± 2 , ± 1 , 0 are shown as polar plots.

l=2

2

l Θ 1(θ θ) l

2

m=1

l=2

2

l Θ 2(θ θ) l

m=2

299

2

l=3 m=0 3

l Θ 0(θ θ) l

2

Figure 77: The angular dependence of the squared modulus of the spherical harmonics | Yml (θ, ϕ) |2 . Since the spherical harmonics only depend on ϕ via the exponential term exp[ i m ϕ ], the probability density is independent of ϕ. The angular dependence is solely determined by the associated Legendre functions Θlm (θ). The θ dependence for l = 3 , m = ± 3 , ± 2 , ± 1 , 0 are shown as polar plots. The probability density is peaked up in the equatorial plane θ = π2 for m = ± l. l=3 m=1

l=3 m=2

l=3

3

l Θ 1(θ θ) l

3

l Θ 2(θ θ) l

3

2

2

l Θ 3(θ θ) l

m=3

300

2

Inversion z

r r θ

y

π−θ π+ϕ

ϕ

x

r -r

Figure 78: In spherical polar coordinates, the inversion or parity operator produces the transformation ϕ → ϕ + π and θ → π − θ. The probability current density is expressed in terms of the wave function Ψ(r, t) as the imaginary part of a vector quantity   ¯h ∗ j(r, t) = =m Ψ (r, t) ∇ Ψ(r, t) (1346) me which in the case of a pure energy eigenvalue is time independent. In spherical polar coordinates, the probability current is expressed as   ∂ ∂ h ¯ Ψ∗ (r) j(r, t) = eˆr Ψ(r) − Ψ(r) Ψ∗ (r) ∂r ∂r 2 me i   ∂ h ¯ ∂ Ψ∗ (r) + eˆθ Ψ(r) − Ψ(r) Ψ∗ (r) ∂θ ∂θ 2 me i r   ∂ h ¯ ∂ Ψ∗ (r) + eˆϕ Ψ(r) − Ψ(r) Ψ∗ (r) ∂ϕ ∂ϕ 2 me i r sin θ (1347) On expressing the wave function as Ψ(r) = g(r)

Θlm (θ)

301

 exp

 imϕ

(1348)

where both g(r) and the associated Legendre functions Θlm (θ) are real, one obtains  2 ¯h m g(r) Θlm (θ) j(r, t) = eˆϕ me r sin θ 2 ¯h m = eˆϕ Ψ(r) (1349) me r sin θ where m is the azimuthal or magnetic quantum number. Hence, if the state is an eigenstate of the z component of the angular momentum, the probability current orbits around the atom in orbits parallel to the x - y plane. ——————————————————————————————————

4.5.14

Exercise 81

In the dipole approximation, the average power radiated when the hydrogen atom decays from the n, l th excited state to the n0 , l0 th excited state is given by Z 2 4 4 e2 3 ∗ Pn,l;n0 ,l0 = ω d r φn0 ,l0 ,m0 (r) r φn,l,m (r) (1350) 0 0 3 n,l;n ,l c3 By considering the angular integration, show that the only non-zero matrix elements occur when the angular momentum quantum numbers l and m for the two states are related via ∆l = l0 − l = ± 1 ∆m = m0 − m = 0 , ± 1

(1351)

These are the selection rules for dipole radiation. From conservation of angular momentum, and since ∆l = ± 1, one sees that the photon must have spin one. The photon has a spin which is an integer multiple of ¯h and, therefore, it is a boson. This contrasts with particles which have spins that are half-odd integer multiples of h ¯ , such as the electron which has 12 ¯h, that are fermions15 . Transitions which do not satisfy the dipole selection rules are forbidden. Generally, forbidden transitions still can occur, but require going beyond the dipole approximation and have intensities which are lower by factors of the order of a2 magnitude λ02 where a0 is the Bohr radius and λ is the wave length of light. —————————————————————————————————— 15 The relationship between spin and statistics was first discussed by W. Pauli in the article, Physical Review, 58, 716 (1940).

302

Energy levels of Hydrogen l=0

0.2

En [ Rydbergs ]

0 -0.2

n=4 n=3 n=2

l=1

l=2

l=3

l=4

Balmer

-0.4

Lyman

-0.6 -0.8 -1

n=1

-1.2

Figure 79: The energy levels for Hydrogen (with 4 ≥ l). The energies are plotted in units of Rydbergs. Some of the transitions allowed by the electric dipole selection rule are shown. 4.5.15

Solution 81

The vector r is expressed in terms of the Cartesian unit vectors through   r = r cos θ eˆz + sin θ ( sin ϕ eˆy + cos ϕ eˆx ) (1352) and substituted into the matrix elements. To evaluate the angular integrations, we start by finding the two recurrence relations l ( 2 l + 1 ) cos θ Pm (cos θ)

=

l−1 l+1 ( l + m ) Pm (cos θ) + ( l − m + 1 ) Pm (cos θ)

l ( 2 l + 1 ) sin θ Pm (cos θ)

l−1 l+1 = Pm+1 (cos θ) + Pm+1 (cos θ)

and then show that cos θ Yml (θ, ϕ) =  1  1 (l − m + 1) (l + m + 1) 2 l+1 (l − m) (l + m) 2 l−1 = Ym + Ym (2l + 1) (2l + 3) (2l − 1) (2l + 1) (1354)   exp + i ϕ sin θ Yml (θ, ϕ) =

303

(1353)

 =

(l + m + 1) (l + m + 2) (2l + 1) (2l + 3)

 12

l+1 Ym+1 +



(l − m) (l − m − 1) (2l − 1) (2l + 1)

 12

l−1 Ym+1

(1355) 



 =

sin θ Yml (θ, ϕ) =

− iϕ

exp

(l − m + 1) (l − m + 2) (2l + 1) (2l + 3)

 12

l+1 Ym−1

 +

(l + m) (l + m − 1) (2l − 1) (2l + 1)

 12

l−1 Ym−1

(1356) The expected result then follows immediately from the orthogonality of the spherical harmonics. ——————————————————————————————————

4.5.16

Exercise 82

Consider the commutation relations ˆ z , x ] = + i ¯h y [L ˆz , y ] = − i h [L ¯ x ˆz , z ] = 0 [L

(1357)

Take the matrix elements of the commutator involving z between states with angular momentum (l0 , m0 ) and (l, m) and use the eigenvalue equation to show that Z 2π Z π 0 ( m − m0 ) dϕ dθ sin θ Yml 0 (θ, ϕ)∗ cos θ Yml (θ, ϕ) = 0 (1358) 0

0

Likewise, by considering the commutation rules involving x and y, show that they can be combined to yield Z 2π Z π 0 ( m − m0 )2 dϕ dθ sin θ Yml 0 (θ, ϕ)∗ sin θ cos ϕ Yml (θ, ϕ) Z =



Z dϕ

0

0 π

0 0

dθ sin θ Yml 0 (θ, ϕ)∗ sin θ cos ϕ Yml (θ, ϕ)

(1359)

0

Hence, deduce the selection rule that the matrix element Z d3 r φ∗n0 ,l0 ,m0 (r) r φn,l,m (r) = 0 vanishes unless m0 − m = ± 1 or m0 − m = 0.

304

(1360)

Following a similar procedure with the double commutator16   ˆ2 , [ L ˆ 2 , r ] ] = 2 ¯h2 r L ˆ2 + L ˆ2 r [L

(1362)

deduce that the matrix elements Z d3 r φ∗n0 ,l0 ,m0 (r) r φn,l,m (r)

(1363)

are only non-zero if l and l0 satisfy the equation    0 2 0 2 (l + l + 1) − 1 (l − l) − 1 = 0

(1364)

Since the first factor is always positive (unless l0 = l = 0), the electric dipole selection rule becomes l0 − l = ± 1. —————————————————————————————————— 4.5.17

Solution 82

The selection rules for the z-component of the electron’s orbital angular momentum can be proved by considering the commutation relations ˆ z , x ] = i ¯h y [L ˆ z , y ] = − i ¯h x [L ˆz , z ] = 0 [L

(1365)

On taking the matrix elements between states with definite z-components of the angular momenta, one finds ˆ z , x ] | nlm > = i ¯h < n0 l0 m0 | y | nlm > < n0 l0 m0 | [ L ˆ z , y ] | mlm > = − i ¯h < n0 l0 m0 | x | nlm > < n0 l0 m0 | [ L ˆ z , z ] | nlm > = 0 < n0 l0 m0 | [ L (1366) which reduce to ( m0 − m ) < n0 l0 m0 | x | nlm > = i < n0 l0 m0 | y | nlm > ( m0 − m ) < n0 l0 m0 | y | mlm > = − i < n0 l0 m0 | x | nlm > ( m0 − m ) < n0 l0 m0 | z | nlm > = 0 (1367) 16 As was shown by Dirac (P. A. M. Dirac, Proc. Roy. Soc. A111, 251 (1926).), this double commutation relation is a specific example of the more general relation

 [ Jˆ2 , [ Jˆ2 , A ] ] = 2 h ¯2

 A Jˆ2 + Jˆ2 A

valid for any arbitrary vector operator A.

305

− 4¯ h2 ( A . Jˆ ) Jˆ

(1361)

From the last equation, it follows that either m0 = m or that < n0 l0 m0 | z | nlm > = 0

(1368)

On combining the first two equations, one finds that ( m0 − m )2 < n0 l0 m0 | x | nlm > = i ( m0 − m ) < n0 l0 m0 | y | nlm > = < n0 l0 m0 | x | nlm > (1369) This equation is solved by requiring that either ( m0 − m )2 = 1

(1370)

< n0 l0 m0 | x | nlm > = 0

(1371)

or Hence, we have derived the selection rules ∆m = ± 1, 0. The selection rules for the magnitude of the electrons orbital angular momentum are found by considering the double commutator   ˆ2 + L ˆ2 r ˆ2 , [ L ˆ 2 , r ] ] = 2 ¯h2 r L [L (1372) On taking the matrix elements of this equation between eigenstates of the magnitude of the orbital angular momentum, one finds ¯4 h



l0 ( l0 + 1 ) − l ( l + 1 )

2

< n0 l0 m0 | r | nlm > =



l0 ( l0 + 1 ) + l ( l + 1 )



(1373) Hence, either < n0 l0 m0 | | nlm > = 0 or



( l0 + l + 1 )2 − 1



( l0 − l )2 − 1

(1374)  = 0

(1375)

Since the first factor is always positive for non-zero l and l0 , the electric dipole selection rule becomes ∆l = ± 1. —————————————————————————————————— 4.5.18

The Addition Theorem

The addition theorem for spherical harmonics states that Pl ( cos γ ) =

4π 2l + 1

mX = l

Yml (θ1 , ϕ1 ) Yml (θ2 , ϕ2 )∗

m = − l

306

(1376)

< n0 l0 m0 | r | nlm >

where γ is the angle between the directions (θ1 , ϕ1 ) and (θ2 , ϕ2 ), i.e., cos γ = cos θ1 cos θ2 + sin θ1 sin θ2 cos ( ϕ1 − ϕ2 ) We shall assume that we can expand Pl ( cos γ ) as a Laplace series X alm1 Yml 1 (θ1 , ϕ1 ) Pl ( cos γ ) =

(1377)

(1378)

m1

where alm1

Z =

dΩ1 Pl ( cos γ ) Yml 1 (θ1 , ϕ1 )

(1379)

But, on noting that the associated Legendre polynomials are the same as the Legendre polynomials when m = 0 and on using eqn(1315), one obtains  Pl ( cos γ ) =

4π 2l + 1

 12

Y0l (γ, 0)

(1380)

The choice of the azimuthal angle as 0 is irrelevant, as m = 0. Hence, we have alm1

 =

4π 2l + 1

 12 Z

dΩ1 Y0l (γ, 0) Yml 1 (θ1 , ϕ1 )∗

(1381)

The choice of the orientation of the axes is irrelevant, as the integration over dΩ1 runs over all 4π solid angle. We shall choose the direction (θ2 , ϕ2 ) as our polar axis. In this case, we need to re-express the spherical harmonic Yml 1 (θ1 , ϕ1 ) in terms of our new variable of integration X (1382) Yml 1 (θ1 , ϕ1 )∗ = blm2 ∗ Yml 2 (γ, 0)∗ m2

where, as we are dealing with an eigenfunction of the scalar magnitude of the angular momentum with eigenvalue governed by l, the eigenvalue l is completely unaltered by a different choice of Cartesian coordinates. Thus, alm1

 =  =  =

4π 2l + 1

 12 X Z

4π 2l + 1

 12 X

4π 2l + 1

 12

dΩγ Y0l (γ, 0) blm2 ∗ Yml 2 (γ, 0)∗

m2

blm2 ∗ δm2 ,0

m2

bl0 ∗

(1383)

where we have used the orthogonality property of the spherical harmonics. Therefore, we see that the coefficients alm which appear in the Laplace expansion are related to only one non-zero coefficient bl0 in their expansion given

307

by eqn(1382). The coefficient bl0 can be obtained directly when γ = 0 as the expansion only contains the one term Yml 1 (θ1 , ϕ1 )∗

= bl0 ∗ Y0l (0, 0)∗ 1  2l + 1 2 l∗ = b0 4π

(1384)

since the spherical harmonics with m 6= 0 vanish when γ = 0. In the above equation the variables (θ1 , ϕ1 ) take on the fixed numerical values (θ2 , ϕ2 ), because γ ≡ 0. Hence,  12  4π bl0 ∗ = Yml 1 (θ2 , ϕ2 )∗ (1385) 2l + 1 Inserting this into eqn(1383), we have   4π l Yml 1 (θ2 , ϕ2 )∗ am1 = 2l + 1

(1386)

On substituting the coefficient alm1 back into the Laplace series expansion, one has m 1 =l X 4π Pl ( cos γ ) = Yml 1 (θ1 , ϕ1 ) Yml 1 (θ2 , ϕ2 )∗ (1387) 2l + 1 m1 =−l

This concludes the proof of the Addition Theorem for Spherical Harmonics.

4.5.19

Finite-Dimensional Representations

ˆ i within One can find a representation of the angular momentum operators L c 2 the manifold of eigenstates of L with the eigenvalue λl . The states are to be represented by column vectors. Since the eigenvalues are ( 2 l + 1 ) fold degenerate, the space is spanned by unit column vectors with ( 2 l + 1 ) rows. Any arbitrary column vector can be expressed as a linear superposition of the ( 2 l + 1 ) basis column vectors. The angular momentum operators act on the states, and transform the states into other states. The rules of transformation correspond to the laws of matrix multiplication, in which the operators are ( 2 l + 1 ) × ( 2 l + 1 ) dimensional square matrices. For convenience, the ˆ z . For definiteness, basis states may be taken to be the ( 2 l + 1 ) eigenstates of L we shall restrict our discussion to l = 1. The three basis eigenstates may be represented by the unit column vectors; ! 1 0 Y11 = (1388) 0 with eigenvalue + ¯h, Y01

=

308

0 1 0

! (1389)

with eigenvalue 0, and 1 Y−1

0 0 1

=

! (1390)

with eigenvalue − ¯ h. Any arbitrary state Ψ with eigenvalue λl can be expressed as the linear superposition of the basis states via Ψ =

m=1 X

Ψm Yml

(1391)

m=−1

in which Ψm are the expansion coefficients. This equation can be expressed more concretely as Ψ1 ! 1 ! 0 ! 0 ! Ψ0 0 1 0 = Ψ1 + Ψ0 + Ψ−1 (1392) Ψ−1 0 0 1 The inner product of two states Ψ and Φ is defined to be a complex number, obtained via m=1 X Φ† Ψ = Φ∗m Ψm (1393) m=−1

The inner product is a generalization of the scalar product to arbitrary dimensional (in this case ( 2 l + 1 ) dimensional) column matrices with complex components. It involves a row vector Φ† which is the complex conjugate and transpose of the column vector Φ. In general, it is convenient to label the components of the basis vectors by an abstract index i, instead of the quantum number m which corresponds to the z component of angular momentum. We shall use this more general labelling below. The inner product is just the result of the matrix multiplication of the row matrix Φ† with the column matrix Ψ, i.e.,   Ψ1 ! † ∗ ∗ ∗ Φ1 Φ2 Φ3 Ψ2 Φ Ψ = (1394) Ψ3 where we are labelling the three components of our states Φ and Ψ by the index i. We shall assume that our states Ψ are normalized to unity, that is, we require †

Ψ Ψ =

3 X

Ψ∗i Ψi = 1

(1395)

i=1

This just corresponds to the condition that the sum over all possibilities yields a probability of unity. The effect of an operator Aˆ on a state Ψ transforms it into another state Φ, according to the laws of matrix multiplication Φ = Aˆ Ψ 309

(1396)

where Aˆ is a three by three matrix. More concretely, the action of the operator is given by ! ! ! Φ1 A1,1 A1,2 A1,3 Ψ1 Φ2 A2,1 A2,2 A2,3 Ψ2 (1397) = Ψ3 Φ3 A3,1 A3,2 A3,3 which results in the state Φ having the components Φi =

j=3 X

Ai,j Ψj

(1398)

j=1

The operator Aˆ is specified by specifying the matrix elements Ai,j . Operators represented by matrices can be compounded by the laws of matrix addition and multiplication. Matrix multiplication is generally non-commutative, and is suitable for representing non-commuting operators. The Hermitean conjugate of the operator is defined by the equation  ∗ † ˆ Φ AΨ = Ψ† Aˆ† Φ (1399) which must be true for any arbitrary states Φ and Ψ. From this, one can see that the Hermitean conjugate of a matrix is just the complex conjugate of the transposed matrix. Thus, if the matrix Aˆ has matrix elements Ai,j , then the Hermitean conjugate Aˆ† has matrix elements given by A∗j,i . A Hermitean matrix is defined to be a matrix for which Aˆ = Aˆ† . ˆ z , one has Since the basis states are eigenfunctions of the operator L ˆ z Ym1 = m ¯h Ym1 L

(1400)

The three linear equations can be trivially interpreted as a matrix equation, in ˆ z is diagonal in the chosen basis. Thus, one can represent which the operator L ˆ z by the diagonal three by three matrix the operator L ˆz = h L ¯

1 0 0

0 0 0 0 0 −1

! (1401)

ˆ z . The effect where the diagonal matrix elements are just the eigenvalues of L ˆ + on the basis states is to raise the angular momentum of the raising operator L according to the law p l ˆ+ Y l = h L ¯ ( l − m ) ( l + m + 1 ) Ym+1 (1402) m ˆ + on the state with maximal m, m = + 1, is to produce Since the effect of L zero, one has one trivial and two non-trivial algebraic equations, which can be 310

used to represent the raising operator as √

0 0 0

ˆ+ = h L ¯

2 √0 ! 2 0 0 0

Likewise, from considering the equation p l ˆ − Yml = h L ¯ ( l + m ) ( l − m + 1 ) Ym−1

(1403)

(1404)

one can construct a representation of the lowering operator as 0 √0 2 √0 0 2

ˆ− = h L ¯

0 ! 0 0

(1405)

The x and y components of the angular momentum are found from the definitions ˆ± = L ˆx ± i L ˆy L (1406) as

ˆx = 1 ( L ˆ+ + L ˆ− ) L 2

and ˆy = L

(1407)

1 ˆ+ − L ˆ− ) (L 2i

(1408)

From which, one has ˆ x = √¯h L 2 and ˆy L

¯ h = √ 2

0 1 0

1 0 1

0 ! 1 0

0 −i 0 i 0 −i 0 i 0

(1409)

! (1410)

It can be shown that these matrices, when compounded according to the laws of matrix multiplication and addition, satisfy the commutation relations X ˆi , L ˆ j ] = i ¯h ˆk i,j,k L (1411) [L k

where i,j,k is the anti-symmetric Levi-Civita symbol. The anti-symmetric LeviCivita symbol is defined so that i,j,k = 1 if i, j, k correspond to an even number of permutations of x, y, z, or is − 1, if i, j, k corresponds to an odd permutation of x, y, z. Otherwise, i,j,k = 0, when one or more index is repeated. Furthermore, since the Hermitean conjugate of the operator just corresponds to the complex conjugate of the transposed matrix, we see that the components ˆ k not only satisfy the angular momentum commutation relations, but are also L

311

Hermitean operators. It can then be easily shown that the operator c2 = L ˆ 2x + L ˆ 2y + L ˆ 2z L

(1412)

is a diagonal matrix, with non-zero matrix elements λl = 1 ( 1 + 1 ) h ¯ 2 . This is as expected, since the basis states are all eigenstates of the magnitude of the angular momentum. Hence, we have 1 0 0 ! c 2 2 0 1 0 L = 2 ¯h (1413) 0

0

1

Since the operator corresponding to the magnitude of the angular momentum is proportional to the unit operator, it commutes with all the components of the ˆ i . The operator L b 2 is sometimes referred to as a Casimir angular momentum L operator. ——————————————————————————————————

4.5.20

Exercise 83

Evaluate the commutators of the three by three matrices representing the components of the angular momentum, and show that they satisfy the commutation relations, such as, ˆx , L ˆ y ] = i ¯h L ˆz [L (1414) Also evaluate the matrix corresponding to the square of the angular momentum c2 and show that it commutes with the various components of L, ˆ and find the L c2 . eigenvalues of L ——————————————————————————————————

4.5.21

Exercise 84

Find the eigenvalues and eigenvectors of the component of the angular momentum along the unit vector ηˆ, where ηˆ = sin θ cos ϕ eˆx + sin θ sin ϕ eˆy + cos θ eˆz

(1415)

ˆ z is diagonal. Work in the space of l = 1 with the basis in which L ˆeˆ(θ) which produces a rotation of the states Hence, find the unitary matrix U through an angle θ, about the axis eˆ = − sin ϕ eˆx + cos ϕ eˆy 312

(1416)

ˆeˆ(θ) should bring the eigenstates of L ˆ z into coincidence with the The operator U ˆ ηˆ. It is also required that U ˆeˆ(θ) does not transform the eigeneigenstates of L ˆ eˆ. states of L ——————————————————————————————————

4.5.22

Solution 84

We seek the eigenvalues and eigenfunctions of the operator ˆ = sin θ cos ϕ L ˆ x + sin θ sin ϕ L ˆ y + cos θ L ˆz η.L which has the matrix representation  cos θ  √1 sin θ exp[ + i ϕ ] ˆ η.L = h ¯  2 0

√1 2

(1417)

 0 √1 sin θ exp[ − i ϕ ]   2 √1 − cos θ 2 (1418) ˆ z is diagonal. The eigenvalues µ h in the basis where L ¯ are determined from the secular equation cos θ − µ sin θ exp[ + i ϕ ] 0

√1 2

sin θ exp[ − i ϕ ] 0 sin θ exp[ + i ϕ ]

sin θ exp[ − i ϕ ] −µ sin θ exp[ + i ϕ ]

0 sin θ exp[ − i ϕ ] √1 − cos θ − µ 2 (1419) and, thus, are found to be µ = 0 , ± ¯h. The eigenfunction Yµη corresponding to the eigenvalue µ, are denoted by  η  αµ Yµη =  βµη  (1420) γµη 0 =

√1 2

√1 2

The components of the µ = 0 eigenfunction are found from the eigenvalue equation √     η  2 cos θ sin θ exp[ − i ϕ ] 0 α0 0  0  =  sin θ exp[ + i ϕ ] 0 sin θ √ exp[ − i ϕ ]   β0η  0 γ0η 0 sin θ exp[ + i ϕ ] − 2 cos θ (1421) Thus, the normalized eigenfunction of the operator in eqn(1417) corresponding to the zero eigenvalue is found as   − sin√ θ exp[ − i ϕ ] 1  (1422) Y0η = √  2 cos θ 2 sin θ exp[ + i ϕ ]

313

Likewise, one finds the eigenfunction corresponding to µ = + 1 is given by   ( 1 + cos √θ ) exp[ − i ϕ ] 1  η  (1423) Y1 = 2 sin θ 2 ( 1 − cos θ ) exp[ + i ϕ ] and the eigenfunction corresponding to µ = − 1 is found as   ( 1 − cos√ θ ) exp[ − i ϕ ] 1  η  Y−1 = − 2 sin θ 2 ( 1 + cos θ ) exp[ + i ϕ ]

(1424)

The phases of the eigenfunctions are completely arbitrary. The required rotation is about an axis in the plane perpendicular to the unit vectors eˆz and eˆη . Since the rotation is specified by three parameters (the axis of the rotation and the angle of rotation), the transformation matrix contains ˆeˆ(θ) is three parameters that have to be determined. The rotation matrix U ˆ partly determined from its action on the eigenstates of Lz ˆeˆ(θ) Yµz Yµη = U

(1425)

There is also the subsidiary requirement that the component of the spins along ˆ e should the rotation axis are unaffected, which means that the eigenstates of L be mapped onto themselves. Thus, we have  η  αµ  βµη  γµη √   ( 1 + cos θ ) exp[ − i ϕ ] − 2 sin θ exp[ − i ϕ ] ( 1 − cos√ θ ) exp[ − i ϕ ] √  = 12  2 sin θ 2 cos θ − 2 sin θ √ ( 1 − cos θ ) exp[ + i ϕ ] 2 sin θ exp[ + i ϕ ] ( 1 + cos θ ) exp[ + i ϕ ]  z  αµ ×  βµz  γµz (1426) ˆe (ϕ) are the eigenvectors that we The columns of the transformation matrix U have just found. The phases of the columns can be altered to fit our requirement that a state with angular momentum directed along the axis of rotation eˆ remains unchanged (apart from an over all phase factor). The phase factors of the Yµη are chosen as exp[ i µ ϕ ] so that the diagonal elements of the unitary transformation are real17 . This choice of phase factors produces the desired ˆ e are unchanged by it. More precisely, under rotation since the eigenstates of L 17 The change of phase can be understood as being the result of a rotation through ϕ about the ηˆ axis

314

ˆ e only change by a θ dependent phase this transformation, the eigenstates of L factor. Also, with the above choice of phase, the eigenstates Yµz are completely unaffected by the transformation, if the angle of rotation θ is set to zero. The above result can be re-written as follows. On recognizing that the component of the angular momentum operator parallel to the axis of rotation is given by     0 − i exp[ − i ϕ ] 0 h ¯ ˆ 0 − i exp[ − i ϕ ]  eˆ . L = √  i exp[ + i ϕ ] 2 0 i exp[ + i ϕ ] 0 (1427) and that the square is given by   2  1 0 − exp[ − 2 i ϕ ] 2 h ¯ ˆ   0 2 0 = eˆ . L 2 − exp[ + 2 i ϕ ] 0 1 we find that the rotation operator may be written in the form    ˆ 2 ˆ  eˆ . L eˆ . L ˆ ˆ Ueˆ(θ) = 1 + cos θ − 1 − i sin θ ¯h ¯h

(1428)

(1429)

This is an example of Cayley’s theorem, as the rotation about the axis eˆ can be written as the exponential of a matrix   i ˆ ˆ Ueˆ(θ) = exp − θ ( eˆ . L ) (1430) ¯h

—————————————————————————————————— The angular momentum operator corresponding to larger l values can also be constructed by the same method, but with a larger number of basis states. In particular, one requires ( 2 l + 1 ) basis states to span the space with total angular momentum l. In fact, almost any operator can be represented by a N × N square matrix acting on a N -dimensional column vector, however, this is only useful for bases where the dimension N is finite.

4.5.23

The Laplacian Operator

The magnitude of the angular momentum can be expressed in terms of the Laplacian by using the vector identity           − A.D B.C A ∧ B . C ∧ D = A.C B.D (1431) 315

which, for our operators, can be adapted as c2 L

= ( r ∧ pˆ ) . ( r ∧ pˆ ) = − ( r ∧ pˆ ) . ( pˆ ∧ rˆ )   = − r . pˆ ( pˆ . r ) − pˆ2 r

(1432)

In manipulating the expression for the square of the angular momentum, one has to carefully respect the ordering of ˆr and p ˆ. In the second line we have used the identity r ∧ pˆ = − pˆ ∧ r (1433) which is valid because the different components of the position and angular momentum commute. The last term in eqn(1432) can be written in terms of r2 p ˆ2 by using the commutator [ r , pˆ2 ] = 2 i ¯h p ˆ

(1434)

Also the first term in eqn(1432) can be re-written by using the commutator pˆ . r − r . pˆ = − 3 i ¯h

(1435)

in the last factor. Hence, we obtain c2 = r2 p ˆ2 + i ¯h r . pˆ − ( r . pˆ ) ( r . pˆ ) L

(1436)

The term r . pˆ is evaluated as r . pˆ = − i ¯h r

∂ ∂r

(1437)

Putting these together, we find that the square of the angular momentum is given by   c 2 ∂ 2 2 ∂ 2 2 r (1438) ˆ + ¯h L = r p ∂r ∂r Thus, we find that the Laplacian operator is expressed in terms of the angular momentum by   c2 L ¯h2 ∂ 2 2 ∂ pˆ = 2 − 2 r (1439) r r ∂r ∂r This results in the usual expression for the kinetic energy in spherical polar coordinates Tˆ

= =

p ˆ2 2m c2 L

¯2 h ∂ − 2 2m 2 m r ∂r

316



∂ r ∂r 2

 (1440)

4.5.24

An Excursion into d-Dimensional Space

As angular momentum is associated with a rotation in a two-dimensional plane and since in d dimensions there is no unique normal to a two-dimensional plane, the angular momentum is not a vector. The angular momentum is a two-form and has d(d − 1)/2 components that are labelled by two indices denoting the ˆ j,k plane of rotation18 . In a Cartesian coordinate system, the components of L are defined via   ∂ ∂ ˆ Lj,k = − i ¯h xj − xk (1441) ∂xk ∂xj ˆ j,k is antisymmetric in the indices j and k. The square of the The quantity L angular momentum no longer has a simple geometric significance, but is still defined as the operator c2 L

=

d X

ˆ 2j,k L

j>k

= −h ¯2

d  X

xj

j>k

∂ ∂ − xk ∂xk ∂xj

2 (1442)

The above form can be expanded as  d d  2 X X ∂ ∂ ∂ 2 2 ∂ c 2 L = − ¯h xj − xj xk + ( d − 1 ) ¯h2 xj 2 ∂xk ∂xj ∂xk ∂x j j=1 j,k

= r2 pˆ2 + ¯h2

d X

xj xk

j,k

∂ ∂ ∂xj ∂xk

+ (d − 1)h ¯2

d X j=1

xj

∂ ∂xj

(1443)

where the radius is defined by the d-dimensional version of the Pythagorean theorem d X r2 = x2j (1444) j=1

A point in a d-dimensional space can be described by hyper-spherical polar coordinates involving the hyper-radius r and (d − 1) angle variables (θ1 , . . . , θd−1 ). The hyper-spherical polar coordinates are related to the Cartesian coordinates via the set of equations xd

= xd−1 = xd−2 = .. .. . . x2 = x1 =

r r r .. . r r

cos θd−1 sin θd−1 cos θd−2 sin θd−1 sin θd−2 cos θd−3

sin θd−1 sin θd−2 . . . sin θ2 cos θ1 sin θd−1 sin θd−2 . . . sin θ2 sin θ1

(1445)

18 In three dimensions, the angular momentum is an antisymmetric second rank tensor that is only masquerading as a vector.

317

The set of equations can be written symbolically as xj = r fj ({θi })

(1446)

Hence, we can evaluated the partial derivative w.r.t. r as   X  ∂xj  ∂ ∂ = ∂r {θi } ∂r {θi } ∂xj j   X ∂ = fj ({θi }) ∂xj j   X xj ∂ = r ∂x j j

(1447)

and the second derivative is evaluated as    2  X  ∂xj  ∂ ∂xk ∂ ∂ = 2 ∂r {θi } ∂r {θi } ∂xj ∂r {θi } ∂xk j,k     X ∂ ∂ = fj ({θi }) fk ({θi }) ∂xj ∂xk j,k

X xj xk ∂ ∂ = 2 r ∂xj ∂xk

(1448)

j,k

Hence, we find that the magnitude of the angular momentum is related to the magnitude the momentum and some radial derivatives 2 ∂ c2 = r2 pˆ2 + ¯h2 r2 ∂ + ( d − 1 ) h L ¯2 r ∂r2 ∂r

(1449)

or pˆ2 = − h ¯2

1 rd−1

∂ ∂r



rd−1

∂ ∂r

 +

c2 L r2

(1450)

This equation can be used to express the kinetic energy in terms of the angular momentum in situations where there is spherical symmetry. The above identity reduces to the usual expression found when d = 3. The eigenvalues of the angular momentum are the hyper-spherical harmonics. Harmonic functions hλµ are defined as homogeneous polynomials of order λ (homogeneous polynomials in the Cartesian coordinates) that satisfy Laplace’s equation ∇2 hλµ = 0 (1451) The hyper-spherical harmonics Yµλ with angular momentum λ can be expressed in terms of the complete set of harmonic functions of order λ via X rλ Yµλ = Cµ,ν hλν (1452) ν

318

where the Cµ,ν are expansion coefficients19 . The above equation implies that the hyper-spherical harmonics only depend on the direction of the vector x. Operating with pˆ2 on rλ Yµλ yields pˆ2 rλ Yµλ

=

0

(1453)

since rλ Yµλ is a combination of harmonic functions. Furthermore, the above equation together with eqn(1450) implies that    c2  1 ∂ ∂ L − ¯ h2 d−1 rd−1 + 2 rλ Yµλ = 0 (1454) r ∂r ∂r r Since the Yµλ only depend on the set of angles {θi } and do not depend on r, we find that the above equation reduces to c2 Y λ = h L ¯ 2 λ ( λ + d − 2 ) Yµλ µ

(1455)

c2 and the eigenThus, the hyper-spherical harmonics Yµλ are eigenfunctions L 2 values are h ¯ λ ( λ + d − 2 ). The eigenvalue equation reduces to the standard eigenvalue equation in three dimensions. ——————————————————————————————————

4.5.25

Exercise 85

Show that the radial momentum pr , which in classical mechanics is defined by 1 (r.p) r must be represented by the symmetrized form   1 1 1 pr = (r.p) + (p.r) 2 r r pr =

in order to yield the Hermitean operator   ∂ 1 pˆr = − i ¯h + ∂r r Show that in d dimensions this leads to the expression   ∂ d − 1 pˆr = − i ¯h + ∂r 2r

(1456)

(1457)

(1458)

(1459)

which is a Hermitean operator. —————————————————————————————————— 19 The expansion coefficients C µ,ν are dependent on the choice of the set of basis functions hλ µ , the choice of the set of quantum numbers µ and an overall phase factor.

319

4.5.26

Solution 85

The quantum operator corresponding to the un-symmetrized form of the classical radial momentum is found from 1 ( r . pˆ ) r

1 ( r eˆr . pˆ ) r   1 = − i ¯h r eˆr . ∇ r   ∂ 1 r = − i ¯h r ∂r ∂ = − i ¯h ∂r =

(1460)

where we have used the expression for ∇ in d-dimensional spherical polar coordinates ∇

= eˆr

∂ 1 ∂ 1 ∂ 1 ∂ + eˆθ1 + eˆθ2 + . . . + eˆθd ∂r r ∂θ1 r sin θ1 ∂θ2 r sin θ1 . . . sin θd−2 ∂θd−1

= eˆr

d−1 X ∂ 1 ∂ + eˆθi ∂r r hi ∂θi i=1

(1461)

In this expression, we have introduced the unit vectors eˆθi in the orthogonal directions defined by 1 ∂ eˆr (1462) eˆθi = hi ∂θi and the product hi of the dimensionless angular factors hi =

i−1 Y



 sin θj

(1463)

j=1

This operator is not Hermitean, as can be seen by constructing the Hermitean conjugate. The Hermitean conjugate is evaluated by considering the effect of the matrix elements with two different arbitrary but normalizable radial wave functions Rn (r) and Rm (r),   Z ∞ ∂ ∗ dr rd−1 Rm (r) − i ¯h Rn (r) (1464) ∂r 0 The factor rd−1 represents the r dependence of the surface area of a d-dimensional sphere, and dr rd−1 is proportional to the volume of an infinitesimal shell of thickness dr. By integration by parts, and using the fact that the wave functions are normalizable to eliminate the boundary terms, one obtains the result   Z ∞ ∂ ∗ = dr Rn (r) + i ¯h rd−1 Rm (r) ∂r 0 320

Z =



dr rd−1 Rn (r)

 + i ¯h

0

∂ d − 1 + i ¯h ∂r r



∗ Rm (r)

(1465) On using the definition of the Hermitean conjugate, one identifies the Hermitean conjugate operator as  †   1 ∂ d − 1 ( r eˆr . pˆ ) = − i ¯h + (1466) r ∂r r

The symmetric partner (p.r) is quantized as   1 pˆ . r r

1 r





= − i ¯h

∇ . r eˆr   = − i ¯h ∇ . eˆr

(1467)

1 r

  d−1 X ∂ 1 ∂ eˆr (1468) = − i ¯h − i ¯h eˆθ . ∂r r hi i ∂θi i=1 The factor of r cancels with 1r in the second line, as the differential operator acts on both factors. In the last line, we have utilized the identity ∂ eˆr = 0 ∂r

(1469)

The derivative of the unit vector eˆr with respect to the angle θi is evaluated as ∂ˆ er = hi eˆθi ∂θi As the unit vectors are orthogonal, one has 1 ( pˆ . r ) r

(1470)

  d−1 X ∂ 1 ∂ = −ih ¯ − i ¯h eˆθ . eˆr ∂r r hi i ∂θi i=1 = − i ¯h

d−1 d−1 X X ∂ 1 − i ¯h eˆθi . hj eˆθj ∂r r hi i=1 j=1

d−1 d−1 X ∂ 1 X = −ih ¯ − i ¯h hj δi−j ∂r r hi j=1 i=1

= − i ¯h

d−1 X ∂ 1 − i ¯h ∂r r i=1

= − i ¯h

∂ d − 1 − i ¯h ∂r r 321

(1471)

Hence, the quantized symmetric partner is equal to the Hermitean conjugate. The Hermitean operator corresponding to the radial component of the momentum is given by the Hermitean part of the operator "  † # 1 1 1 pˆr = ( r . pˆ ) + ( r . pˆ ) 2 r r   1 1 1 ( r . pˆ ) + ( pˆ . r ) = 2 r r   ∂ d − 1 = − i ¯h + (1472) ∂r 2r

——————————————————————————————————

322

4.6

Spherically Symmetric Potentials

The Hamiltonian for a particle of mass m and charge q moving in a spherically symmetric potential is given by p ˆ2 + q φ(r) (1473) 2m where the potential φ(r) only depends on the radial coordinate r and is independent of (θ, ϕ). Since the Hamiltonian commutes with the orbital angular momentum operator, ˆ , L ˆ] = 0 [H (1474) ˆ = H

and, therefore, also commutes with the magnitude c2 ] = 0 ˆ , L [H

(1475)

c2 . These eigenfunctions ˆ Lˆz and L one may find simultaneous eigenfunctions of H, must be of the separable form ΨE,l,m (r) = RE,l (r) Yml (θ, ϕ)

(1476)

Thus, the energy eigenvalue equation becomes     c2 ∂ L h2 ¯ 2 ∂ + q φ(r) ΨE,l,m (r) = E ΨE,l,m (r) − r + 2 m r2 2 m r2 ∂r ∂r (1477) After substituting eqn(1476) in the above equation, and after using the fact that Yml (θ, ϕ) is an eigenstate of L2 , with eigenvalue ¯h2 l ( l + 1 ), one finds that the spherical harmonic is a common factor that can be cancelled. The radial part of the wave function must satisfy the radial equation     h2 ¯ ∂ ¯h2 l ( l + 1 ) 2 ∂ − r + + q φ(r) RE,l (r) = E RE,l (r) 2 m r2 ∂r ∂r 2 m r2 (1478) This is an ordinary differential equation, instead of a partial differential equation, in which the centrifugal potential appears as an addition to the electrostatic potential. The radial equation is independent of the azimuthal quantum number m, and hence, so are the radial eigenfunctions and eigenvalues. ——————————————————————————————————

4.6.1

Exercise 86

A particle of mass m moves in a three-dimensional spherically symmetric potential V (r) which vanishes at r → ∞. It has an eigenstate which is given by     Φ(r) = C rs exp

− αr

323

sinl θ exp

ilϕ

(1479)

Effective Radial Potential 0.4

Centrifugal Barrier 2

Veff(r)

0.2

Veff(r) = V(r) + ! l ( l + 1 ) / 2 m r

2

0 0

5

10

15

20

l=0

-0.2

l=1 l=2 -0.4

r

Figure 80: The effective radial potential Vef f (r) is the sum of the potential V (r) and the centrifugal potential. What are the values of the angular momentum and energy quantum numbers for this state? Find the potential V (r). ——————————————————————————————————

4.6.2

Solution 86

The state is an eigenstate of angular momentum with l = l and m = l. The energy eigenvalue and the potential can be determined from   ¯h2 (1480) ∇2 Φ(r) V (r) − E Φ(r) = 2m where ∇2 Φ(r)

     1 ∂2 l(l + 1) l s r − C r exp − α r sin θ exp i l ϕ r ∂r2 r2   (1 + s) s(s + 1) − l(l + 1) Φ(r) = α2 − 2 α + r r2 (1481) 

=

324

Thus, at r → ∞ where V (r) → 0, we find E = −

¯ 2 α2 h 2m

(1482)

and ¯2 h 2m

V (r) =



(s − l)(s + l + 1) 2α(1 + s) − 2 r r

 (1483)

——————————————————————————————————

4.6.3

The Free Particle

The free particle corresponds to the case of a vanishing electrostatic potential φ(r) = 0. The time-independent Schr¨odinger equation for the free particle separates in spherical polar coordinates. The radial wave function is given by the solution of     ∂ ¯h2 l ( l + 1 ) h2 ¯ 2 ∂ r + RE,l (r) = E RE,l (r) − 2 m r2 ∂r ∂r 2 m r2 (1484) √ which, with h ¯ k = 2 m E, becomes     ∂ l(l + 1) 1 ∂ r2 + RE,l (r) = k 2 RE,l (r) (1485) − 2 r ∂r ∂r r2 It is also useful to introduce a dimensionless variable ρ = k r to obtain the scaled equation     l(l + 1) 1 ∂ 2 ∂ − ρ + RE,l (ρ/k) = RE,l (ρ/k) (1486) ρ2 ∂ρ ∂ρ ρ2 This differential equation reduces to Bessel’s equation. This can be seen by making the transformation RE,l (r) =

yl (ρ) √ ρ

(1487)

which leads to d 2 yl 1 d yl + + 2 dρ ρ dρ



( l + 12 )2 1 − ρ2

 yl = 0

(1488)

This is Bessel’s equation for half-integer indices, and the solutions we seek are proportional to the Bessel functions Jl+ 12 (ρ). The Bessel functions are finite at 325

the origin ρ = 0. The solutions for RE,l are, therefore, known as the spherical Bessel functions r π Jl+ 12 (ρ) RE,l (r) = jl (ρ) = (1489) √ 2 ρ Equation (1486) actually has two independent solutions jl (z) and ηl (z) which are the spherical Bessel functions and the spherical Neumann functions. For values of ρ close to the origin, the solutions have leading terms in the power series expansion which vary as Rl (r) ∝ ρα + O(ρα+2 )

(1490)

where α has to be determined. On substituting the leading term into the differential equation, and noting that the term of order unity can be neglected compared with the centrifugal barrier near r = 0, one finds that the differential equation reduces to   − α ( α + 1 ) + l ( l + 1 ) ρα−2 + O(ρα ) = 0 (1491) for small ρ. The values of α, for which this is satisfied, satisfy a quadratic equation and are given by either of the solutions α = l or α = − ( l + 1 ). The spherical Bessel functions jl (ρ) vary as ρl close to the origin, which is a consequence of the high centrifugal barrier that prevents classical particles with fixed l approaching the origin. The spherical Neumann functions ηl (ρ) vary as ρ−(l+1) as ρ → 0. The spherical Neumann functions are to be discarded as r2 | RE,l (r) |2 dr is proportional to the probability of finding a particle in the interval dr at a radial distance r. The inclusion of the spherical Neumann functions would lead to the probability diverging like r−2l as r → 0. Therefore, the spherical Neumann functions represent unphysical solutions near the origin. Since the spherical Bessel functions and spherical Neumann functions play an important role in the description of scattering by short ranged potentials, we shall digress to examine the general properties of these functions in more detail. The solution of equation(1486) with α = l, is the spherical Bessel function jl (z), and can be written as jl (z) = D z l

Z

+1

ds exp[ i s z ] ( 1 − s2 )l

(1492)

−1

where the usual normalization is given as Dl =

1 2l+1 l!

(1493)

It can be shown that jl (z) satisfies the differential equation for Rl,E by direct substitution. We first need to evaluate the various derivatives in the equation. 326

The first derivative of jl (z) is given by ∂ jl (z) ∂z

+1

Z

= D l z l−1

ds exp[ i s z ] ( 1 − s2 )l

−1 +1

+ D i zl

Z

ds exp[ i s z ] s ( 1 − s2 )l

(1494)

−1

and the second derivative is given by ∂2 jl (z) ∂z 2

= D l ( l − 1 ) z l−2

Z

+1

ds exp[ i s z ] ( 1 − s2 )l

−1

+ D 2 i l z l−1

Z

+1

ds exp[ i s z ] s ( 1 − s2 )l

−1

Z

− D zl

+1

ds exp[ i s z ] s2 ( 1 − s2 )l

(1495)

−1

On substituting in the differential equation number (1486) and collecting like terms, after some cancellations, we find   2 ∂ l(l + 1) ∂2 jl (z) + jl (z) + 1 − jl (z) ∂z 2 z ∂z z2 Z +1 = D 2 i ( l + 1 ) z l−1 ds exp[ i s z ] s ( 1 − s2 )l −1

+ D zl

Z

+1

ds exp[ i s z ] ( 1 − s2 )l+1

−1

= −Diz

l−1

Z

+1

ds exp[ i s z ] −1

+ D zl

Z

d ( 1 − s2 )l+1 ds

+1

ds exp[ i s z ] ( 1 − s2 )l+1

(1496)

−1

In the last expression, we have rewritten the integrand in the first term as a differential with respect to s. On integrating by parts in the first term, and noting the boundary term is zero, we find that the right hand side of the differential equation is given by Z +1 = − D zl ds exp[ i s z ] ( 1 − s2 )l+1 −1 +1

+ D zl

Z

ds exp[ i s z ] ( 1 − s2 )l+1

−1

=

0

(1497)

Thus, jl (z) satisfies the radial differential equation.

327

From the above integral representation expression for, jl and its derivative, we can easily verify that the recursion relation jl+1 (z) =

∂ l jl (z) − jl (z) z ∂z

(1498)

holds true. A second useful recursion relation that can be derived is (2l + 1)

∂ jl (z) = l jl−1 (z) − ( l + 1 ) jl+1 (z) ∂z

(1499)

A second solution ηl (z) is given by the integral ηl (z) = C

1 z l+1

Z

+1

ds exp[ i s z ] −1

1 ( 1 − s2 )l+1

(1500)

since the differential equation is invariant under the transformation l → − ( l + 1 ). From the integral expressions for jl (z) and ηl (z), one can show that the spherical Bessel functions are regular at the origin lim jl (z) ∼

z → 0

2l l! zl ( 2 l + 1 )!

(1501)

while the spherical Neumann functions are divergent at the origin, lim ηl (z) ∼ −

z → 0

( 2 l )! −(l+1) z 2l l!

(1502)

The asymptotic large z variation of the spherical Bessel function is found from the integral representation Z +1 zl jl (z) = l+1 ds exp[ i s z ] ( 1 − s2 )l (1503) 2 l! −1 by integrating by parts l times. Each integration by parts reduces the power of the mononomial in z in front of the integral by unity, yielding jl (z) =

il 2l+1 l!

Z

+1

 ds exp[ i s z ]

−1

d ds

l

( 1 − s2 )l

(1504)

since the boundary terms all vanish. The vanishing of the boundary terms occurs as there are more factors of ( 1 − s2 ) than there are derivatives with respect to s. However, on further integration by parts, the higher order derivatives remain non-zero at the boundaries s = ± 1. Repeated integration by parts yields an 328

expansion where the non-zero terms are ordered in increasing powers of z −1 . The first non-zero term in the asymptotic expansion is given by il+1 jl (z) ∼ − l+1 exp[ i s z ] 2 l! z



d ds

l

s=+1 ( 1 − s ) + O(z −2 ) (1505) 2

l

s=−1

Thus, the leading term in the asymptotic large z expansion is given by lim

z → ∞

jl (z) ∼

cos[ z − ( l + 1 ) z

π 2

]

(1506)

The asymptotic expansion for the spherical Neumann function is given by an analogous expression which is most easily found from eqn(1506) via the substitution l → − ( l + 1 ), and is given by lim

z → ∞

ηl (z) ∼

sin[ z − ( l + 1 ) z

π 2

]

(1507)

The explicit forms of the first few spherical Bessel and Neumann functions are given in Table(5). The eigenfunctions for the free particle have to be square integrable over any finite volume of space, as the integral of the modulus squared of the wave function represents the probability that a particle is found in the volume. Since only the spherical Bessel functions are integrable over the region near the origin, they can represent physical wave functions near the origin. The spherical Neumann functions have to be discarded near the origin. The orthogonality condition for the spherical Bessel functions is given by Z ∞ π δ( k − k 0 ) (1508) dr r2 jl (kr) jl (k 0 r) = 2 k2 0 The normalization condition corresponds to the continuous eigenvalue spectrum associated with the delocalized wave functions. As the wave functions are proportional to the spherical harmonics, they are orthogonal for different values of l. Therefore, we only require that the Bessel functions with the same values of l but different energy eigenvalues are orthogonal. The solution of the free particle energy eigenvalue equation, with angular momentum l, also has uses in cases where there is a spherically symmetric potential φ(r) which is short ranged. In this case, the energy eigenvalue equation can be separated into radial and angular parts Ψ(r, θ, ϕ) = Rl,k (r) Yml (θ, ϕ)

(1509)

since the potential is spherically symmetric. The wave functions of the spatially extended states with energy eigenvalues Ek > 0 have asymptotic large r 329

Table 5: The Spherical Bessel jl (z) and Neumann ηl (z) Functions. l

jl (z)

l=0

j0 (z)

sin z z

l=1

j1 (z)

sin z − z cos z z2

l=2

j2 (z)

( 3 − z 2 ) sin z − 3 z cos z z3

l=3

j3 (z)

( 15 − 6 z 2 ) sin z − ( 15 z − z 3 ) cos z z4

l

ηl (z)

l=0

η0 (z)

l=1

η1 (z)

l=2

η2 (z)

l=3

η3 (z)

− − − −

cos z z

cos z + z sin z z2

( 3 − z 2 ) cos z + 3 z sin z z3

( 15 − 6 z 2 ) cos z + ( 15 z − z 3 ) sin z z4

330

Spherical Bessel Functions 1.2

j0(x)

1 0.8 0.6

j1(x) j2(x)

0.4 0.2 0 0

2

4

6

8

10

12

-0.2 -0.4

Figure 81: The spherical Bessel functions jl (z) as a function of z for various l. limits which are expressible as linear superpositions of the spherical Bessel and Neumann functions, since φ(r) ≈ 0 in this region. Thus, Rl,k (r) = Al jl (kr) + Bl ηl (kr)

(1510)

The ratio of the amplitude of the two terms contains information about the short ranged potential, and is usually expressed in terms of the phase shift δl (k), Bl = − tan δl (k) Al

(1511)

On using this definition of the phase shift, and the asymptotic large r forms of the spherical Bessel and Neumann functions, one finds that the asymptotic form of the radial wave function is given by   sin k r − l 2π + δl (k) Al Rl,k (r) ∼ (1512) cos δl (k) kr Hence, the large r form of the radial wave function resembles that of a free particle. The main difference is that the phase of the radial oscillations is shifted by δl (k). The normalization of the extended states, for an arbitrary short ranged potential, can be inferred from the asymptotic large r behavior. This can be 331

Spherical Neumann Functions 0.4

η0(x) η1(x)

0.3 0.2

η2(x)

0.1 0 0

2

4

6

8

10

12

-0.1 -0.2 -0.3 -0.4

x

Figure 82: The spherical Neumann functions ηl (z) as a function of z for various l. The spherical Neumann functions diverge at the origin. shown by re-writing radial part of the energy eigenvalue equation in the alternate form     1 ∂2 2mE l(l + 1) 2 m q φ(r) r R (r) + R (r) = + Rl,k (r) l,k l,k r ∂r2 r2 h2 ¯ ¯h2 (1513) A generalized continuity equation can be derived from the time-independent Schr¨ odinger equation. The continuity equation has the form ( k 2 − k 02 )

Z

R ∗ dr r2 Rl,k 0 (r) Rl,k (r) =

0

 = −

r

∗ Rl,k 0 (r)

∂ ∂ ∗ r Rl,k (r) − r Rl,k r Rl,k 0 ∂r ∂r

R (1514) 0

We note that as r → 0, Rl,k (r) vanishes for non-zero l. Hence, the normalization is governed solely by the asymptotic behavior. As previously mentioned, the radial wave function has the asymptotic large r form   1 lπ Rl,k (r) ∼ sin k r − + δl (k) (1515) r 2 where δl (k) is the phase shift and k 2 = 332

2 m E . h ¯2

Then, on substituting the

asymptotic form in the expression for the normalization, one finds Z R ∗ dr r2 Rl,k 0 (r) Rl,k (r) = 0      1 lπ lπ 0 0 k sin k R − + δ (k ) cos k R − + δ (k) = 02 l l k − k2 2 2     lπ lπ − k 0 sin k R − + δl (k) cos k 0 R − + δl (k 0 ) 2 2 (1516) The above expression can be re-written in the form   " sin (k 0 − k) R + δ (k 0 ) − δ (k) l l π = 2 ( k0 − k )   0 0 sin (k + k) R + δl (k ) + δl (k) − lπ # + ( k0 + k ) (1517) The region of physical interest is that in which both k and k 0 are both positive. In the limit r → ∞, the first term dominates the above expression and so the normalization integral can be expressed as π = δ( k 0 − k ) (1518) 2 where we have used the representation of the delta function sin N x → δ(x) (1519) N →∞ x This shows that the orthonormality condition of the continuum states must involve the Dirac delta function Z ∞ π ∗ dr r2 Rl,k δ( k 0 − k ) (1520) 0 (r) Rl,k (r) = 2 0 lim

The othonormality condition of the spherical Bessel functions provides one concrete example of this relation. For the free particle, the radial wave function is identified with the spherical Bessel function by Rl (r) = k jl (kr)

(1521)

since we require that, when r → ∞, the radial wave function has the same normalization as the asymptotic form given in eqn(1515). Therefore, one recovers the identity Z ∞ π dr r2 jl (k 0 r) jl (kr) = δ( k 0 − k ) (1522) 2 2 k 0 where k and k 0 are both restricted to have positive values.

333

The Dirac delta Function 70 60

N

sin N x / π x

50

200

40

100 30

50

20

25

10 0 -10 -20 -0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

x

Figure 83: The Dirac delta function δ(x) can be regarded as the limit N → ∞ of the sequence of functions sinπNx x . 4.6.4

The Spherical Square Well

The spherically symmetric potential represents a short ranged potential, in which q φ(r) = − V0 for r < a and q φ(r) = 0 for r > a. The states of angular momentum l have radial wave functions which satisfy   h2 ¯ d ¯h2 l ( l + 1 ) 2 d R − r + R = ( E + V0 ) R (1523) 2 m r2 dr dr 2 m r2 for r < a and −

¯h2 d 2 2 m r dr



r2

dR dr

 +

¯2 l ( l + 1 ) h R = ER 2 m r2

(1524)

for r > a. The condition for regularity at the origin leads us to select the solution R(r) = Cl jl (k 0 r) (1525) for r < a, where k 0 is given by ¯h2 k 02 = 2 m ( E + V0 )

(1526)

If E > 0, the solutions outside the well are oscillatory and correspond to linear combinations of the spherical Bessel and spherical Neumann functions. These 334

solutions correspond to scattering states. On the other hand, the bound states have energies E < 0 which correspond to imaginary values of k. The bound state radial wave functions ought to vanish as r → ∞. Scattering States Inside the well, the radial wave function with angular momentum l must be proportional to the spherical Bessel function with wave vector k 0 Rl,E (r) = Cl jl (k 0 r)

(1527)

for r < a. As previously mentioned, the spherical Neumann function diverges at the origin and, therefore, is not an physically acceptable solution inside the well. Outside the well, r > a, the scattering states can be written as the linear combination Rl,E (r) = Al jl (kr) + Bl ηl (kr) (1528) The ratio of the coefficients Al and Bl can be obtained from the continuity conditions at r = a. These conditions are that the radial wave function is continuous at r = a, lim Rl,E (a + ) =

 → 0

lim Rl,E (a − )

 → 0

and that the derivative is also continuous at r = a ∂ ∂ Rl,E (r) = lim Rl,E (r) lim  → 0 ∂r  → 0 ∂r a+ a−

(1529)

(1530)

These equations can be divided, yielding Al jl0 (ka) + Bl ηl0 (ka) k 0 jl0 (k 0 a) = k jl (k 0 a) Al jl (ka) + Bl ηl (ka)

(1531)

Thus, the ratio of Bl to Al is determined as Bl k 0 jl0 (k 0 a) jl (ka) − k jl0 (ka) jl (k 0 a) = Al k ηl0 (ka) jl (k 0 a) − k 0 jl0 (k 0 a) ηl (ka)

(1532)

This ratio is usually written in terms of the tangent of the phase shift δl (k). The phase shift (modulo 2 π) can be inferred from experiments in which a beam of particles of energy E falls incident on the potential. For low energies, the scattering is obviously dominated by the l = 0 phase shift, since classically the angular momentum is given by l = k r. Hence, only the l = 0 contribution is expected to be important at low energies where k → 0. For the case l = 0, the matching procedure leads to B0 k 0 cos k 0 a sin ka − k cos ka sin k 0 a = 0 A0 k cos k 0 a cos ka + k sin ka sin k 0 a 335

(1533)

The ratio of amplitudes can be expressed in terms of the l = 0 phase shift δ0 (k), via the definition B0 − tan δ0 (k) = (1534) A0 Hence, the phase shift is given by  k  ( k0 ) tan k 0 a − tan k a tan δ0 (k) = 1 + ( kk0 ) tan k 0 a tan k a or δ0 (k) = tan−1



k tan k 0 a k0

(1535)

 − ka

(1536)

where ¯h2 k 02 = ¯h2 k 2 + 2 m V0 . The phase shift vanishes identically for k 0 = k and, therefore, is a non-linear measure of the strength of the potential. 0 The phase shift δ0 (k) (in units of π) is shown in fig(84) as a function of kπa for fixed small values of the energy. The phase shifts are only slightly dependent on k at low energies. It is seen that the phase shift rapidly increases by π at various

Phase Shift for the attractive spherical potential well 6

5

π δ0(k)/π

4

3

2

1

0 0

1

2

3

4

5

6

π k'a/π

Figure 84: The l = 0 phase shift δ0 (k) for low-energy particles incident on an attractive potential well, as a function of k 0 a. values of k 0 a. These values are tabulated in Table(6). At these particular values of k 0 a, the potential is almost strong enough to produce another bound state with l = 0. This can be demonstrated by noting that these special values of k 0 a correspond to k 0 a = (n + 12 ) π, so the radial wave function inside the potential 336

0

Table 6: The values of ( kπa ) near which the l = 0 phase shift changes by π and resonant scattering occurs. 



l = 0 bound state

k0 a π

n=1 n=2 n=3 n=4 n=5 n=6

0.48 1.50 2.51 3.50 4.51 5.51

well has a node at the origin and has zero slope at the edge of the well. The wave function with zero slope inside the well only matches to a wave function with zero slope outside the well. Outside the well, the eigenfunction is constant (since k = 0) and corresponds to having a zero value of the decay constant κ and, therefore, represents a zero energy bound state. As seen in fig(85), the scattering cross-sections are large for low-energy particles incident on potentials where V0 is close to these critical values. This type of scattering is known as

dσ/dΩ

Scattering cross-section

0

1

2

3

4

5

6

k'a/π

Figure 85: The l = 0 scattering cross-section for low-energy particles incident on an attractive potential well, as a function of k 0 a.

337

resonant scattering. In the limit k → 0, the matching condition reduces to   tan k 0 a B0 = ka 1 − A0 k0 a

(1537)

Hence, the ratio vanishes proportionally to k, as k → 0. This means that the asymptotic, large r, radial wave function has the form of   B0 cos kr sin kr lim R0,k (r) = A0 − k →0 kr A0 kr   asc (1538) = A0 1 − r where

 asc = a

1 −

tan k 0 a k0 a

 (1539)

is the scattering length20 . In the limit of low energies, the differential scattering cross-section is given by dσ = | asc |2 (1540) dΩ which is isotropic, that is, independent of the scattering angle θ. —————————————————————————————————— Example: l = 0 Scattering from a Repulsive Spherical Well Potential. In this case, we consider the energy eigenstates with l = 0 of the spherically symmetric potential V (r) = V0 V (r) = 0

if r < a if r > a

(1541)

We shall consider the case where E < V0 . By analytic continuation from the case where V0 > E, the radial wave function on the inside of the well has the form sinh κ r R0 (r) = C0 (1542) κr which is proportional to the l = 0 spherical Bessel function for imaginary argument. This function satisfies the radial equation, if the energy eigenvalue is related to the potential V0 and κ via E = V0 −

¯h2 κ2 2m

(1543)

20 Note that if the magnitude of the potential is such that k 0 a is close to an odd multiple of π2 , the scattering length can be anomalously large. When this occurs, low energy particles are resonantly scattered.

338

Outside the well, the radial wave function is given by a linear combination of the l = 0 spherical Bessel and Neumann functions R0 (r) = A0

cos k r sin k r − B0 kr kr

(1544)

where k is related to the energy eigenvalue by E =

¯ 2 k2 h 2m

(1545)

The matching conditions for the radial wave function at r = a lead to the equation tanh κ a tan( k a + δ0 (k) ) = (1546) ka κa The matching condition leads to the phase shift being given by   k tanh κ a − ka (1547) δ0 (k) = tan−1 κ In the limit of a hard sphere, V0 → ∞, so κ a → ∞. In the hard sphere limit, the above expression for the phase shift simplifies to lim

V0 →∞

δ0 (k) → − k a

(1548)

The k → 0 limit of the scattering amplitude is found as asc

= − lim

k→0

= a

tan δ0 (k) k (1549)

Therefore, at low energies where the scattering is isotropic and independent of k, the total scattering cross-section of the hard sphere is found as   Z dσ σT = dΩ dΩ Z = dΩ a2sc = 4 π a2

(1550)

This derivation is only valid for small k. The total cross-section should be compared with geometrical cross-section, which is defined as the area of (the front half of) the target’s surface projected onto the plane of the beam. As seen in fig(86), the geometrical cross-section for the hard sphere is given by π a2 , the area of a disk of radius a. Therefore, as the total scattering cross-section of the hard sphere potential is 4 π a2 , the total cross-section is four times greater than the geometrical cross-section.

339

Geometric cross-section

Target Incident Beam Wave Front

Figure 86: The geometrical scattering cross-section for particles incident on a hard sphere. —————————————————————————————————— Bound States In order to determine the bound states it is necessary to examine the behavior of the radial wave function for imaginary wave vectors as r → ∞. It is useful to define certain combinations of the spherical Bessel and Neumann functions that prove to be useful. From the asymptotic variation of the Bessel and Neumann functions, one recognizes that the combinations (1)

hl (ρ) = jl (ρ) + i ηl (ρ) (2)

hl (ρ) = jl (ρ) − i ηl (ρ)

(1551)

known as the spherical Hankel functions of the first and second kind, have simple exponential variations    1 π (1) hl (ρ) = exp + i ρ − (l + 1) ρ 2    1 π (2) hl (ρ) = exp − i ρ − (l + 1) (1552) ρ 2 For imaginary k, such that ρ = + i κ r, only the spherical Hankel functions of the first kind decay exponentially at r → ∞. Thus, when discussing bound 340

(m)

Table 7: The spherical Hankel Functions hl (1)

(2)

hl (z) −

l=0 −i

l=1

−i

l=2 l=3

−i

i z

hl (z)

exp[ + i z ]

(1−iz) z2

+

exp[ + i z ]

(3−3iz−z 2 ) z3

i

exp[ + i z ]

(15−15iz−6z 2 +iz 3 ) z4

(z)

i

exp[ + i z ]

i

i z

exp[ − i z ]

(1+iz) z2

exp[ − i z ]

(3+3iz−z 2 ) z3

exp[ − i z ]

(15+15iz−6z 2 −iz 3 ) z4

exp[ − i z ]

states it is useful to express the radial wave function outside the well in terms of the spherical Hankel function of the first kind. The wave functions in the interior and exterior regions should be matched at r = a. For the bound states, since E < 0, the radial wave functions exterior (1) to the well have the form hl (i κ r) while the wave functions interior to the 0 well have the form jl (k r). Therefore, on equating the logarithmic derivatives, one finds the condition 0

(1)

i κ hl (i κ a) k 0 jl0 (k 0 a) = 0 (1) jl (k a) hl (i κ a)

(1553)

which, together with the relation k 02 −

2 m V0 = − κ2 ¯h2

(1554)

leads to discrete values of κ and, hence, discrete energy eigenvalues Eκ,l . This procedure for graphically determining the allowed values of κ illustrated for the l = 0 bound states in fig(87). The dashed curves shows the relationship between κ and k 0 of eqn(1554), for different values of V0 . The intersection of the solid lines and the dashed lines yield the allowed values of κ for the l = 0 bound states. The number of intersections and the bound states increase as V0 increases. The wave function for the lowest energy l = 0 bound state is shown in fig(88), for various values of V0 . ——————————————————————————————————

341

Graphical Solution for Bound State Energies 8

κa

- cot k'a / k' a

6

4

2

0 0

0.5

1

1.5

2

-2

-4

k' a / π

Figure 87: The graphical solution of the l = 0 version of the self consistency condition (eqn(1553)), for various values of V0 . 4.6.5

Exercise 87

The deuteron consists of a neutron which is bound in an l = 0 state to a proton by a short-ranged attractive potential. The neutron and proton in the deuteron are in a spin triplet state. Assume that the potential can be represented by a spherically symmetric “square” well of depth V0 and radius a. The experimentally determined binding energy E0 is 2.21 MeV. The scattering length atsc for a neutron-proton system in a triplet spin state is inferred, from low-energy elastic scattering experiments21 , to have a magnitude of 5.38 F. Using this data and the reduced mass, estimate the depth of the potential V0 and the radius a. The estimated value of a can be compared with the value of 2.1 F for the rms radius of the charge distribution deuteron and 0.8 F for the rms radius for the charge distribution of the proton, obtained by the diffraction of high-energy electrons22 . The neutron-proton interaction is highly spin-dependent. For example, the experimentally determined singlet neutron-proton scattering length is given by assc = − 23.7 F. —————————————————————————————————— 21 L.J.

Rainwater, W.W. Havens, J.R. Dunning and C.S. Wu, Phys. Rev. 73, 733 (1948). Chambers and R. Hofstadter, Phys. Rev. 103, 1454 (1956), and J.A. McIntyre and R. Hofstadter, Phys. Rev. 98, 158 (1955). 22 E.E.

342

The radial wave function for the l=0 bound state 1.5

rR0(r)

1

0.5

0 0

0.5

1

1.5

2

r/a

Figure 88: The lowest energy l = 0 bound state wave function, for various values of V0 . 4.6.6

Solution 87

The triplet scattering length atsc = 5.38 F, was measured in experiments involving the scattering of very low-energy particles. For the model potential, the scattering length is given by   tan k 00 a asc = a 1 − (1555) k 00 a where the wave vector k 00 is associated with a zero-energy particle within the well. Therefore, ¯h2 k 002 = V0 (1556) 2µ If the potential only has one bound state and the scattering is resonant, then the product k 00 a must be slightly greater than π2 . The bound state energy E0 is related to the imaginary wave vector κ via E0 = − where µ =

m 2

¯ 2 κ2 h 2µ

(1557)

is the reduced mass of the neutron-proton system. The bound

343

state energy is determined from the condition k 0 cot k 0 a = − κ where V0 =

¯2 h 2µ



k 02 + κ2

(1558)  (1559)

or k 002 = k 02 + κ2

(1560)

There are two parameters that need to be determined a and k 00 . We shall eliminate a in terms of k 0 , via √   κ k 002 − κ2 −1 κa = √ tan − (1561) κ k 002 − κ2 where the inverse tangent has to be understood as giving an angle greater than π k00 2 . The equation involves the dimensionless variable x = κ . The resulting value of a is to be substituted in the equation for the scattering length, yielding the equation   √ −1 2  p   x tan − x − 1  1 1 √ κ atsc = √ tan−1 − x2 − 1 − tan x x2 − 1 x2 − 1 (1562) This equation involves the dimensionless product κ atsc which for triplet scattering has a value κ atsc ∼ 1.23. This leads to the value of x = 4.05. The depth of the potential is then given by V0 = E0 x2

(1563)

which is estimated to be V0t = 36.25 MeV. The width of the well is given by a = 2.01 F. Since the experimentally determined singlet neutron-proton scattering length is given by assc = − 23.7 F, then using the value of a just calculated, one determines that the depth of the singlet potential is V0s = 23.7 MeV. The singlet potential is not deep enough to bind the neutron and proton. Similar results were obtained by using a more realistic form of the potential23 . —————————————————————————————————— 23 P.M.

Morse, J.B. Fisk and L.I. Schiff, Phys. Rev. 50, 748 (1936).

344

4.6.7

Exercise 88

Find the energy eigenvalues of the l = 0 states of the spherically symmetric infinite square well potential, ∞

if r > a

0

if r < a

V (r) = (1564)

——————————————————————————————————

4.6.8

Exercise 89

Show that the spherical square well has no bound states unless V0 a2 >

¯ 2 π2 h 8m

(1565)

——————————————————————————————————

4.6.9

Solution 89

The bound state solution outside the well has to have the form of a decaying exponential and, therefore, is given by a spherical Hankel function in which k → i κ, and is given by a spherical Bessel function inside the well, with wave vector k 0 , where ¯h2 k 2 ¯h2 κ2 − V0 = − (1566) 2m 2m On matching the exponentially decaying spherical Hankel functions with imaginary k with the spherical Bessel functions of argument k 0 r at r = a one obtains a transcendental equation for the bound state. The largest value of κ occurs, for fixed V0 , when l = 0 as this minimizes the effect of the rotational kinetic energy or centrifugal potential and, therefore, maximizes the effect of the attractive potential. For l = 0, the bound state condition simplifies to k 0 cot k 0 a = − κ

(1567)

On squaring this equation, one obtains k 02 cot2 k 0 a = + κ2 =

345

2 m V0 − k 02 ¯h2

(1568)

which contains some spurious solutions. The left hand side is greater or equal to zero, and only falls to zero for k 0 equal to every odd multiple of 2πa . The right hand side is an inverse parabola, which becomes negative for k 02 greater V0 than 2 m . The solutions of the squared equation are, therefore, restricted to h ¯2 the region of relatively small k 0 where the inverted parabola is positive. The solutions of the squared equation correspond to the solutions of the original equation only in the restricted range where cot k 0 a is negative. Thus, the lowest energy bound states occur for 2πa < k 0 < πa . Hence, on increasing the strength of V0 from zero, the lowest energy bound state first occurs when π2 2 m V0 ≥ 4 a2 ¯h2

(1569)

——————————————————————————————————

4.6.10

Exercise 90

Show that if a square well just binds an energy level of angular momentum l, the parameters satisfy the equation  r 2 m V0 a2 jl−1 = 0 (1570) ¯h2 which utilized the recurrence formulas for the Bessel functions. ——————————————————————————————————

4.6.11

Solution 90

Inside the well of radius a, a > r, the solution for the radial wave function, with angular momentum l, is proportional to the spherical Bessel function Rl (r) = Al jl (kr)

(1571)

The wave vector k is related to the energy eigenvalue by E =

¯ 2 k2 h − V0 2m

(1572)

Since the spherical Neumann function ηl (kr) diverges at r = 0, it yields an unphysical probability density and, therefore, is discarded for r < a. Outside the well, r > a, the solution is proportional to the spherical Hankel function (1) Rl (r) = Bl hl (iκr) (1573) 346

where (1)

hl (iκr) = jl (iκr) + i ηl (iκr)

(1574)

Since E ≤ 0, the momentum is imaginary and is given by E = −

¯ 2 κ2 h 2m

(1575)

(1)

The spherical Hankel function hl (iκr) represents the combination of the possible solutions which decays exponentially as r → ∞. The other combination (2) hl (iκr) blows up as r → ∞ and so is not normalizable and, therefore, the unphysical solution is discarded. The bound state wave functions must satisfy the continuity conditions at r = a, which can be combined to yield 0(1)

i κ hl (iκa) k jl0 (ka) = (1) jl (ka) hl (iκa)

(1576)

If the potential is so weak that it just binds the particle, then E ∼ 0, and hence κ = 0. The spherical Hankel functions are dominated by the spherical Neumann functions as they diverge like ( κ a )−(l+1) in the limit κ → 0. Hence, in the limit κ → 0, the matching condition becomes (l + 1) k jl0 (ka) = − jl (ka) a

(1577)

k a jl0 (ka) + ( l + 1 ) jl (ka) = 0

(1578)

and we have the condition

The recurrence relations for the spherical Bessel functions can be combined to yield ( l + 1 ) jl (ka) + ka jl0 (ka) = ka jl−1 (ka) (1579) Hence, on using the above recurrence relation, we find that the condition that the spherical square potential well is just sufficiently strong that it starts have a bound state of angular momentum l is given by jl−1 (ka) = 0

(1580)

in which

2 m V0 ¯h2 since E = 0. Thus, we have derived the required condition. k2 =

(1581)

——————————————————————————————————

347

4.6.12

Exercise 91

Find the bound states corresponding to angular momentum l = 0 for the attractive delta function spherical shell potential V (r) = − V0 a δ( | r | − a )

(1582)

Find the minimum value of V0 for which a bound state occurs. ——————————————————————————————————

4.6.13

Solution 91

The energy eigenfunction function for a spherically symmetric Hamiltonian can be separated in terms of a radial function R(r) and an angular function Yml (θ, ϕ), Φ(r) = Rn,l (r) Yml (θ, ϕ)

(1583)

The solution for the radial function inside the potential shell is given by the analytic continuation of the spherical Bessel function, of angular momentum l, Rn,l (r) = Al jl ( i κ r )

for

r < a

(1584)

since the wave function must be normalizable at the origin. However, outside the shell it is given by a linear combination of the spherical Bessel and Neumann functions that decay as r → ∞ exponentially. This combination defines the spherical Hankel functions, Rn,l (r) = Bl jl ( i κ r ) + Cl ηl ( i κ r )

for

r > a

(1585)

The wave function is continuous at r = a. Thus, Rn,l (a + ) = Rn,l (a − )

(1586)

The delta function only depends on r and so the first derivative has a discontinuous slope at r = a, given by integrating the radial equation  2  Z a+ h ¯ 1 ∂ 2 ∂ dr r2 r + V a δ( r − a ) + E Rn,l (r) = 0 (1587) 0 2 m r2 ∂r ∂r a− Taking the limit  → 0, one obtains 0 0 Rn,l (a + ) − Rn,l (a − ) +

2 m V0 a R(a) = 0 ¯h2

(1588)

The energy eigenvalue is given by E = − 348

¯ 2 κ2 h 2m

(1589)

For l = 0, one has Rn,0 (r)

= A0

sinh κ r r exp

Rn,0 (r)

for 

r < a

− κr

= B0

for r > a r On eliminating B and A, we find the eigenvalue equation for α   2 m V0 a exp κ a κ = sinh κ a ¯h2

(1590)

(1591)

The bound state just forms when the solution κ → 0, where sinh κ a → κ a. In this case we have ¯h2 V0 = (1592) 2 m a2 ——————————————————————————————————

4.6.14

Exercise 92

A particle is confined to move in the volume between two concentric spheres of radius R0 and radius R1 , where R1 > R0 . The potential is radially symmetric and has the form V (r) = 0 V (r) → ∞

if

R1 > r > R0 elsewhere

(1593)

(a) Find the values of the energy eigenvalues if the particle has angular momentum given by l = 0. (b) Consider the limit R1 → R0 , which resembles a particle rigidly bound at radius R1 , but is free to rotate. Assume that the particle is in the ground state. The excitation energy of any energy state is defined as the difference between its energy eigenvalue and the energy of the ground state. In this limit, show that the excitation energies for fixed l are independent of the value of l. Also, evaluate all of the remaining finite excitation energies. ——————————————————————————————————

4.6.15

Solution 92

The general solution for the energy eigenstates of angular momentum l is of the form   φl (r) =

Al jl (kr) + Bl ηl (kr) 349

Yml (θ, ϕ)

(1594)

for R1 > r > R0 , and zero elsewhere. The wave function has to satisfy the boundary conditions   Al jl (kR1 ) + Bl ηl (kR1 ) = 0 (1595) at r = R1 and



 Al jl (kR0 ) + Bl ηl (kR0 )

= 0

(1596)

at r = R0 . These equations can be used to eliminate the ratio of a to B, yielding a transcendental equation for k jl (kR1 ) jl (kR0 ) = ηl (kR0 ) ηl (kR1 )

(1597)

For l = 0, this can be re-written as sin k ( R0 − R1 ) = 0

(1598)

which yields the allowed values of k as kn =

nπ R1 − R0

(1599)

and has the energy eigenvalue En,l=0 =

¯ 2 n2 π 2 h 2 m ( R1 − R0 )2

(1600)

In the limit R1 → R0 , the energy differences between the eigenstates of different n (with fixed l) are independent of l as the radial equation becomes     ∂ ¯h2 l ( l + 1 ) h2 ¯ 2 ∂Rl (r) − r R (r) = E − Rl (r) (1601) l 2 m R02 ∂r ∂r 2 m R02 Thus, the differences in the eigenvalues, for fixed l, are identical to the differences with l = 0, which we have shown tend to infinity. The only remaining finite excitation energies correspond to states of the same n values but with different rotational energies. That is we have a rigid rotor, with excitation energies   ¯h2 0 0 El0 −l = l ( l + 1 ) − l ( l + 1 ) (1602) 2 m R02 This corresponds to the excitation energies of a three-dimensional rigid rotor, with moment of inertia I = m R02 . The difference between successive excita2 tion energies increase with increasing l as h¯I ( l + 1 ). These differences in excitation energies are frequently observed in the optical absorption spectra of molecules. ——————————————————————————————————

350

4.6.16

Ladder operators for a free particle

The radial wave function Rl (r) for a free particle with angular momentum l satisfies the eigenvalue equation      2  2 ∂Rl l(l + 1) ∂ Rl + − 1 Rl 0 = − − ∂ρ2 ρ ∂ρ ρ2     1 ∂ 2 ρRl l(l + 1) 0 = − + − 1 Rl (1603) ρ ∂ρ2 ρ2 where the dimensionless variable ρ is defined by ρ = kr

(1604)

and k is related to the energy eigenvalue E via E =

¯ 2 k2 h 2m

(1605)

We shall introduce ladder operators that transform radial wave functions Rl (r) into radial wave functions with the same value of the energy but with different values of the angular momentum 24 . Since the ladder operators that are to be introduced keep the energy E unchanged, the raising and lowering operators can be expressed in terms of the variable ρ and the derivative with respect to ρ. The raising operator Aˆ+ l is defined via 1 ∂ l Aˆ+ ρ + l = − ρ ∂ρ ρ

(1606)

The lowering operator Aˆ− l is the Hermitean conjugate of the raising operator Aˆ+ . In three dimensions, the lowering operator Aˆ− l l is calculated as l 1 ∂ Aˆ− ρ + l = + ρ ∂ρ ρ

(1607)

The products of the raising and lowering operators are given by 1 ∂2 l(l + 1) ˆ− Aˆ+ ρ + l Al = − ρ ∂ρ2 ρ2

(1608)

or, when taken in the opposite order 1 ∂2 l(l − 1) ˆ+ Aˆ− ρ + l Al = − ρ ∂ρ2 ρ2

(1609)

On comparing the above two forms, we find the relation ˆ+ ˆ+ ˆ− Aˆ− l Al = Al−1 Al−1 24 L.

Infeld, Phys. Rev. 59, 737 (1941).

351

(1610)

Radial distribution function 1.5

1

2

jl (ρ) ρ) ρ

2

l-0 l=1 l=2 l=3

0.5

0 0

1

2

3

4

5

6

ρ/π

Figure 89: The (un-normalized) radial distribution for a free particle with angular momentum l. The centrifugal barrier excludes particles with finite values of l from a region surrounding the origin. Thus, we have found two equivalent forms of the operator that appears in the radial equation with angular momentum (l − 1). When acting on Rl , the lowering operator Aˆ− l produces the radial wave function Rl−1 which is an energy eigenstate with the same energy. This can be seen by first using eqn(1608) to write the radial equation in the form   + ˆ− ˆ Al Al − 1 Rl = 0 (1611) and then acting on this with Aˆ− l . This yields the equation   − + ˆ− ˆ ˆ Al Al Al − 1 Rl = 0 which can be re-written as 

ˆ+ Aˆ− l Al − 1



Aˆ− l Rl = 0

Hence, on using eqn(1610), one finds   − + ˆ ˆ Al−1 Al−1 − 1 Aˆ− l Rl = 0 352

(1612)

(1613)

(1614)

which indicates that Aˆ− l Rl is proportional to the radial wave function Rl−1 with angular momentum (l − 1) and the same energy E Rl−1 ∝ Aˆ− l Rl

(1615)

Since the angular momentum is reduced by unity, the above relation justifies naming Aˆ− l as the lowering operator. The raising operator Aˆ+ l+1 acting on Rl produces a radial wave function Rl+1 where the angular momentum is raised by one unit. This can be shown by first substituting l by (l + 1) in eqn(1610), to produce ˆ+ ˆ+ ˆ− Aˆ− l+1 Al+1 = Al Al The radial equation for Rl can be written as   ˆ+ − 1 Rl = 0 Aˆ− A l+1 l+1

(1616)

(1617)

If the raising operator Aˆ+ l+1 acts on the radial equation as expressed in eqn(1617), one has   + − + ˆ ˆ ˆ 0 = Al+1 Al+1 Al+1 − 1 Rl   − + ˆ ˆ (1618) 0 = Al+1 Al+1 − 1 Aˆ+ l+1 Rl The factor in the parentheses is identified as the radial equation operator with angular momentum (l + 1), therefore the product Aˆ+ l+1 Rl must be the radial wave function with angular momentum (l + 1). Hence, we have Rl+1 ∝ Aˆ+ l+1 Rl

(1619)

so the raising operator Aˆ+ l+1 increases the angular momentum of the radial wave function Rl by one unit, but keeps the energy unchanged. The radial wave function with l = 0 can be determined from the radial equation, which we write in the form  2  ∂ ρR0 = − ρR0 (1620) ∂ρ2 The general solution is easily found and is given by ρ R0 (r) = A0 sin ρ + B0 cos ρ

(1621)

The physically acceptable solution has B0 = 0 since the probability of finding the particle in the neighborhood of ρ → 0 must be finite. Hence, the radial wave function with l = 0 is given by   sin ρ R0 (r) = A0 (1622) ρ 353

which is proportional to the zero-th order spherical Bessel function j0 (ρ) which is given by   sin ρ j0 (ρ) = (1623) ρ

The radial wave functions with l 6= 0 can be found by successive use of the raising operators. For example, the radial wave function with l = 1 is found from j0 (ρ) by   1 ∂ 1 R1 (r) ∝ − ρ + R0 (r) ρ ∂ρ ρ   1 sin ρ 1 ∂ sin ρ + A0 = − A1 ρ ∂ρ ρ ρ   sin ρ − ρ cos ρ = A1 (1624) ρ2 The radial wave function R1 (r) is proportional to the spherical Bessel function j1 (ρ), which is given by   sin ρ − ρ cos ρ (1625) j1 (ρ) = ρ2 It is seen that the spherical Bessel function of order one tends to zero like ρ when ρ → 0. Repeated applications of the raising operators yield the un-normalized radial functions in the form Rl (r) = Al jl (ρ)

(1626)

where the coefficient Al is an arbitrary complex number. These solutions, like the spherical Bessel functions, are all well-behaved at the origin. Furthermore, the probability of finding the particle near the origin must decrease as l increases, due to the increase of the centrifugal barrier in the effective potential. The spherical Neumann functions ηl (ρ) of order l are independent solutions of the radial equation with angular momentum l. However, these solutions are singular at the origin, as can be seen by examining the spherical Neumann function of order zero   cos ρ η0 (ρ) = − (1627) ρ which diverges like ρ−1 when ρ → 0. Other independent solutions of the radial equation can be obtained by successive actions of the raising operators. For example, the spherical Neumann function of order one is obtained from η0 (ρ) by   1 ∂ 1 η1 (ρ) = − ρ + η0 (ρ) ρ ∂ρ ρ 354



1 ∂ 1 ρ + ρ ∂ρ ρ   cos ρ + ρ sin ρ = − ρ2

= −





cos ρ ρ



(1628)

The function η1 (ρ) diverges like ρ−2 as ρ → 0. The general form of ηl (ρ) as ρ → 0 can be determined by induction. One can show that if lim ηl (ρ) ∼ −

ρ→0

then since

 −

ηl+1 (ρ) =

1

(1629)

ρl+1

1 ∂ (l + 1) ρ + ρ ∂ρ ρ

 ηl (ρ)

(1630)

one must have  lim ηl+1 (ρ) ∼



ρ→0

= −



(l + 1) 1 ∂ 1 + ρ ∂ρ ρl ρl+2



1

(1631)

ρl+2

Since the asymptotic form holds true for l = 0, one has shown that lim ηl (ρ) ∼ −

ρ→0

1 ρl+1

(1632)

is true for all values of l. The asymptotic ρ → 0 behavior of the spherical Bessel functions is given by lim jl (ρ) ∼ ρl (1633) ρ→0

The above two asymptotic forms of the independent solutions can be simply obtained from the radial equation, if one assumes that the leading terms of the solutions are of the form R ∼ ρα for some unknown values of α.

4.6.17

The Rayleigh Equation

A plane wave can be expanded in a series of spherical waves, via the Rayleigh equation   X exp i k r cos θ = il ( 2 l + 1 ) jl ( kr ) Pl ( cos θ ) (1634) l

The expansion can be thought of in terms as the decomposition of a wave front (a plane of constant phase) of a beam in terms of its angular momentum components, as is indicated schematically in fig(90). The above expansion is to be expected since the eigenstates of the Laplacian with eigenvalue − k 2 , which are simultaneous eigenfunctions of the linear momentum must be related to 355

Classical decomposition of a beam into angular momentum eigenstates l=4

l=k r l=3 l=2 l=1 l=0

Cross section of the Beam

Figure 90: A semi-classical description of the decomposition of a plane wave in terms of states with angular momentum l. If a point on the wave front at a distance r from the axis with ( l + 12 ) ¯h > p r > ( l − 21 ) ¯h has angular momentum l. the eigenstates of the Laplacian which are simultaneous eigenfunctions of the angular momentum. The Rayleigh expansion can be verified by differentiating the above relation with respect to k r leading to   X i cos θ exp i k r cos θ = il ( 2 l + 1 ) jl0 ( kr ) Pl ( cos θ ) l

cos θ

X

l+1

i

( 2 l + 1 ) jl ( kr ) Pl ( cos θ ) =

l

X

il ( 2 l + 1 ) jl0 ( kr ) Pl ( cos θ )

l

(1635) The recurrence relation ( 2 l + 1 ) z Pl (z) = ( l + 1 ) Pl+1 (z) + l Pl−1 (z)

(1636)

can be used to write X

l+1

i

 jl ( kr )

 ( l + 1 ) Pl+1 ( cos θ ) + l Pl−1 ( cos θ )

l

=

X

il ( 2 l + 1 ) jl0 ( kr ) Pl ( cos θ )

l

356

(1637)

The recurrence relation for jl0 ( kr ) given by ( 2 l + 1 ) jl0 ( kr ) = l jl−1 ( kr ) − ( l + 1 ) jl+1 ( kr )

(1638)

can be used in eqn(1637) to obtain   X il+1 jl ( kr ) ( l + 1 ) Pl+1 ( cos θ ) + l Pl−1 ( cos θ ) l

=

X

il



 l jl−1 ( kr ) − ( l + 1 ) jl+1 ( kr )

Pl ( cos θ )

(1639)

l

which, on changing the index from l + 1 to l in the first term and l − 1 to l in the second term, yields the equality  X  l i jl−1 ( kr ) l Pl ( cos θ ) − jl+1 ( kr ) ( l + 1 ) Pl ( cos θ ) l

=

X

l

i



 l jl−1 ( kr ) − ( l + 1 ) jl+1 ( kr )

Pl ( cos θ )

(1640)

l

This concludes the proof of the Rayleigh expansion. An alternate proof can be found by using the orthogonality of the Legendre polynomials and forming the matrix elements with the plane wave expansion   X exp i k r cos θ = al0 Pl0 ( cos θ ) (1641) l0

with unknown expansion coefficients al0 . Then, on multiplying by Pl (cos θ) sin θ and integrating w.r.t θ, and using the orthogonality condition, one finds   Z 1 X 2 δl,l0 d cos θ Pl ( cos θ ) exp i k r cos θ = al0 (1642) 2 l + 1 −1 0 l

Thus, the expansion coefficients are evaluated as   Z 1 (2l + 1) al = ds exp i s k r Pl (s) 2 −1 =

( 2 l + 1 ) il jl ( kr )

(1643)

where we have used the Rodriguez formula expression for the Legendre polynomial 1 ∂l ( s2 − 1 )l (1644) Pl (s) = l 2 l! ∂sl and integrated by parts l times and then used an integral representation of the spherical Bessel function   Z +1 zl jl (z) = l+1 ds exp i z s ( 1 − s2 )l (1645) 2 l! −1 357

The above two expressions allowed us to express the integral involving the Legendre polynomial in terms of the spherical Bessel function. The Rayleigh expansion is often used in considerations of scattering.

4.6.18

The Isotropic Planar Harmonic Oscillator

The potential for the Isotropic Planar Harmonic Oscillator is given by V (r) =

m ω2 2 r 2

(1646)

which only depends on the radial distance r. In planar polar coordinates (r, ϕ), the energy eigenvalue equation for the isotropic two-dimensional Harmonic Oscillator is of the form " # ∂ ¯h2 ∂2 m ω 2 r2 ¯h2 1 ∂ r − + − E Ψ(r, ϕ) = 0 (1647) − 2 m r ∂r ∂r 2 m r2 ∂ϕ2 2 The Laplacian in the eigenvalue equation is expressed in the form appropriate for two-dimensional motion. In particular, it can be seen that this operator is Hermitean if one uses the two-dimensional measure of “volume”, dr r dϕ, in the definition of the inner product. The potential only depends on the radial ˆ z comdistance r, and so the two-dimensional angular momentum operator L mutes with the Hamiltonian. Thus, the energy and angular momentum are ˆ and L ˆz. conserved. Also, it is possible to find simultaneous eigenfunctions of H ˆ The eigenvalue of Lz is denoted by h ¯ mz . On introducing the separable form for the eigenfunction   1 √ Ψ(r, ϕ) = R(r) exp i mz ϕ (1648) 2π the differential equation for the radial wave function becomes " # h2 1 ∂ ¯ ∂ ¯h2 m2z m ω 2 r2 − r + + − E R(r) = 0 2 m r ∂r ∂r 2 m r2 2

(1649)

On introducing the dimensionless variable, ρ, defined by ρ =

m ω r2 ¯h

(1650)

it is found that the derivative with respect to r is given in terms of the differential with respect to ρ, via ∂ ∂r

= =

∂ρ ∂ ∂r ∂ρ r 1 mω ∂ 2 2ρ ¯h ∂ρ 358

(1651)

Thus, the radial differential equation takes the dimensionless form   ∂2 ∂ m2z ρ E hω − 2ρ ¯ − 2 + + − R(r) = 0 ∂ρ2 ∂ρ 2ρ 2 ¯h ω Furthermore, the radial function is parameterize in the form   ρ | m2z | R(r) = ρ exp − L(ρ) 2

(1652)

(1653)

which has the physically acceptable asymptotic variation at the boundaries, mz ρ → 0 and ρ → ∞. The term proportional to ρ| 2 | indicates that we seek the solution which is normalizable at ρ = 0. The exponential factor indicates that we seek the solution which decays exponentially as ρ → ∞. The first order derivative of R(r) is evaluated as !   ∂R ρ ∂L |mz | 1 | m2z | = ρ exp − + L − L (1654) ∂ρ 2 ∂ρ 2ρ 2 and the second order derivative is given by a lengthier, but analogous, expression. The differential equation becomes   ∂L E ∂2L + 2 ( ρ − |mz | − 1 ) + |mz | + 1 − L = 0 (1655) −2ρ ∂ρ2 ∂ρ ¯h ω which is the differential equation for the associated Laguerre polynomials. The differential equation can be solved by the Fr¨obenius method, in which we expand the solution in powers of ρ L(ρ) =

∞ X

an ρn

(1656)

n=0

On substituting the series into the differential equation, and demanding that the coefficient of ρn should vanish, one finds the recurrence relation   1 E ( n + 1 ) ( n + |mz | + 1 ) an+1 = n + ( |mz | + 1 − ) an (1657) 2 ¯h ω which relates an+1 to an . Thus, all the coefficients in the series are determined in terms of a0 . The value of a0 is only determined from the normalization of the wave function. If the series did not truncate, then one can show that the solution would diverge exponentially as ρ → ∞, and would not be normalizable. Thus, in order to satisfy the boundary conditions the series must truncate. The series truncates at the nr -th term, so anr +1 = anr +2 = anr +3 = . . . = 0, if the energy eigenvalue is given by E = h ¯ ω ( 2 nr + |mz | + 1 )

359

(1658)

In Cartesian coordinates, the energy eigenvalue equation separates into two ordinary differential equations, and each differential equation represents a onedimensional Harmonic Oscillator. The energy eigenvalue corresponds to the sum of the energy eigenvalues for the two one-dimensional problems. Therefore, the energy is given by 1 1 + ny + ) 2 2

E = h ¯ ω ( nx +

(1659)

which contains a contribution of h ¯ ω from the energy of the zero-point motion.

4.6.19

The Spherical Harmonic Oscillator

The potential for the spherically symmetric Harmonic Oscillator is given by V (r) =

m ω2 2 r 2

(1660)

b2 As the potential is only a function of r, the angular momentum operators L ˆ z commute with the Hamiltonian. Thus, angular momentum is conserved, and L ˆ L b 2 and L ˆz. and one can find simultaneous eigenstates of the operators H, The energy eigenvalue equation for the Spherically Symmetric Harmonic Oscillator is separable in spherical polar coordinates. If the wave function is assumed to be of the form Ψ(r, θ, ϕ) = Rl (r) Yml (θ, ϕ)

(1661)

then, after using the properties of the spherical harmonics, it is found that the radial wave function R(r) satisfies an equation independent of the variables, θ and ϕ. The radial equation for the spherically symmetric harmonic oscillator is given by    2  d ¯h l ( l + 1 ) m ω 2 r2 h2 ¯ 2 d R − r + + R = ER 2 m r2 dr dr 2 m r2 2 (1662) The solution can be written in the form   mω 2 l Rl (r) = r exp − r f (r) 2 ¯h

(1663)

which minimizes the effect of the centrifugal potential barrier at the origin, and the exponential term takes care of the binding at asymptotic large values of r. It should be noted that a second solution exists which, at large r, varies asymptotically as   mω 2 R(r) ∼ exp + r (1664) 2 ¯h 360

This solution has been discarded as it is not normalizable, and hence, unphysical. On changing variable to m ω r2 ρ = (1665) ¯h and on substituting the wave function into eqn(1662), one finds     3 ∂f 1 E 3 ∂2f ρ + + l − ρ + − ( + l ) f = 0 (1666) ∂ρ2 2 ∂ρ 2 ¯h ω 2 q h ¯ρ which leads to f ( mω ) being recognized as being proportional to an associated Laguerre polynomial in the variable ρ. The associated Laguerre polynomial Lca (z) is a solution of the differential equation z

∂2 ∂ L(z) + ( c + 1 − z ) L(z) + a L(z) = 0 ∂z 2 ∂z

(1667)

The solutions of this equation only satisfy the appropriate boundary conditions at z → ∞ for specific values of a. This discrete set of a values give rise to the discrete eigenvalues of the energy E of the harmonic oscillator. This can be seen by substituting the Fr¨ obenius series for L(z) X L(z) = an z n (1668) n

into the radial equation and combining like powers of z n . The form of L(z) satisfies the differential equation if the coefficients of each power of z vanish identically. This procedure results in the recursion relations for the coefficients an , ( n + 1 ) ( n + c + 1 ) an+1 = ( n − a ) an (1669) If the series did not terminate, then the asymptotic large n behavior would result in the sum of the series L(ρ) having a positive exponential variation (L(ρ) ∼ exp[ + ρ ]) since an ∼ an−1 n . If this hypothetical form of the exponential factor for the series sum was combined with the explicit exponential factor which we extracted from the radial wave function via the substitution   l ρ R(ρ) = ρ 2 exp − L(ρ) (1670) 2 this would lead to a solution which has the second form of asymptotic variation (R(ρ) ∼ exp[ + ρ2 ]) as r → ∞. If this were the case, the radial wave function would not be square integrable and would not represent a bound state. Therefore, the series must terminate. For the series to terminate, one requires that n = a for some n. For an integer value of a, say equal to nr , the Laguerre polynomials truncate after the first ( nr + 1 ) terms, and all the higher order coefficients vanish, anr +1 = anr +2 = . . . = 0. The series expansion always converges and the wave function satisfies appropriate boundary conditions at r → 0 and at r → ∞. For the isotropic three-dimensional harmonic oscillator, 361

enforcing the requirement that the boundary conditions are satisfied, i.e. the series truncates, yields the eigenvalues as   3 E = h ¯ ω + l + 2 nr (1671) 2 The factor of 23 is the sum of the three independent zero-point motions, and 2 nr + l are the number of excited quanta. Since each state with given l has a degeneracy of 2 l + 1, and states with N quanta have a degeneracy corresponding to the range of values of l from 0 to N , the states with N quanta are highly degenerate. The degeneracy of a state with energy E = h ¯ ω ( 32 + N ) is equal to ( N + 2 )2( N + 1 ) .

Radial wave functions for the spherical harmonic oscillator 3

nr=0,1,2 l=0

Rn,l

2

1s 2s 3s

1

0 0

1

2

3

4

-1

ρ1/2

Figure 91: The radial wavefunction Rn,l (r) for the simple harmonic oscillator for l = 0 and nr = 0 , 1 , 2. The radial quantum number nr determines the number of nodes in the radial wave function. ——————————————————————————————————

4.6.20

Exercise 93

Show that the substitution of l



R(r) = ρ 2 exp

362



ρ 2

 L(ρ)

(1672)

Radial wave functions for the spherical harmonic oscillator 2

nr=0 l=0,1,2

1.5

Rnl

1s 1p

1

1d 0.5

0 0

1

2

3

4

1/2

ρ

Figure 92: The radial wavefunction Rn,l (r) for the simple harmonic oscillator for nr = 0 and l = 0 , 1 , 2. For non-zero values of l, the centrifugal barrier prevents particles from reaching the origin. where ρ =

mω 2 r ¯h

(1673)

into the radial equation −

¯ 2 1 ∂ 2 ∂R h ¯2 l ( l + 1 ) h m ω2 2 r + R + r R = ER 2 2 2 m r ∂r ∂r 2mr 2

(1674)

leads to the differential equation for the associated Laguerre polynomials. ——————————————————————————————————

4.6.21

Solution 93

On defining the dimensionless variable ρ through ρ =

mω 2 r ¯h

one finds that ∂ ∂ρ ∂ = = 2 ∂r ∂r ∂ρ

363

r

mω 1 ∂ ρ2 ¯h ∂ρ

(1675)

(1676)

The radial differential equation becomes " # hω ¯ 4 1 ∂ 3 ∂ l(l + 1) 2E 0 = − ρ2 ρ2 − − ρ + R 2 ρ ∂ρ ∂ρ ρ ¯h ω " # hω ¯ ∂2 ∂ l(l + 1) 2E = − 4ρ + 6 − − ρ + R 2 ∂ρ2 ∂ρ ρ ¯h ω (1677) The derivatives of l



R(r) = ρ 2 exp



ρ 2

 L(ρ)

(1678)

are evaluated as       l ρ ∂ 1 l ρ l l −1 ρ ∂R L(ρ) L(ρ) + ρ 2 exp − L(ρ) − ρ 2 exp − = ρ2 exp − 2 ∂ρ 2 2 2 ∂ρ 2 (1679) and     l ∂2R l l ρ l l −1 ρ −2 2 2 = ( − 1)ρ L(ρ) − ρ L(ρ) exp − exp − ∂ρ2 2 2 2 4 2   l l ρ ∂ + ρ 2 −1 exp − L(ρ) 2 2 ∂ρ     1 l ρ ρ l l −1 2 2 L(ρ) L(ρ) + ρ exp − exp − − ρ 2 2 4 4   1 l ρ ∂ − ρ 2 exp − L(ρ) 2 2 ∂ρ     l l −1 ρ ∂ 1 l ρ ∂ 2 2 + ρ exp − L(ρ) − ρ exp − L(ρ) 2 2 ∂ρ 2 2 ∂ρ   2 l ρ ∂ + ρ 2 exp − L(ρ) (1680) 2 ∂ρ2 Hence, we have    l ∂2R ρ ∂2 ∂ l(l − 2) 2 exp = ρ − 4 ρ + 4 ( l − ρ ) + + ρ − 2 l L(ρ) ∂ρ2 2 ∂ρ2 ∂ρ ρ (1681) and    l ∂R ρ ∂ 3l 6 = ρ 2 exp − 6 − 3 + L(ρ) (1682) ∂ρ 2 ∂ρ ρ 4ρ

which on substituting into the radial equation leads to  "  hω l ¯ ρ ∂2 ∂ 2E − ρ 2 exp − 4ρ 2 +(4l+6−4ρ) + − ( 2 l + 3 ) L(ρ) = 0 2 2 ∂ρ ∂ρ ¯h ω (1683) 364

On cancelling the common factor and dividing by 4, one obtains "  ∂2 3 ∂ E 1 ρ + (l + − ρ) + − ( 2 l + 3 ) L(ρ) = 0 (1684) ∂ρ2 2 ∂ρ 2 ¯h ω 4 which is the associated Laguerre equation, as was to be shown. ——————————————————————————————————

4.6.22

Exercise 94

Consider the set of energy eigenstates of the isotropic harmonic oscillator with E 3 5 7 h ¯ ω = 2 , 2 and 2 . Show that the eigenfunctions of the lowest energy eigenstates are degenerate and that the wave functions and degeneracy found in the spherical polar representation coincide with the representation in terms of three independent one-dimensional harmonic oscillators. Evaluate the degeneracy of the energy eigenvalue with N quanta present. —————————————————————————————————— The symmetry group of the spherical harmonic oscillator. The degeneracy of a spherical harmonic oscillator can be discussed in terms of operators which, when acting on one state, produce other states degenerate with it. These operators only have finite matrix elements between degenerate states and must commute with the Hamiltonian. The commutators of these operators must also commute with the Hamiltonian and, therefore, the operators should generate a Lie algebra. Since the angular momentum must commute with the Hamiltonian, the Lie algebra of the three-dimensional harmonic oscillator should be greater than, and include, the angular momentum Lie algebra. The Hamiltonian of the three-dimensional harmonic oscillator can be written in terms of the creation and annihilation operators of the three one-dimensional harmonic oscillators via 3 ¯ω X ˆ = h ( a†i ai + ai a†i ) H 2 i=1

(1685)

Using the expressions for the position and momentum operators in terms of the creation and annihilation operators    12  ¯h † aj + aj (1686) x ˆj = 2mω and  pˆk = i

¯h m ω 2

 12 

365

a†k

 − ak

(1687)

one can show that the operators representing the components of the angular momentum are given by X ˆ i = − i ¯h L i,j,k a†j ak (1688) j,k

where i,j,k is the anti-symmetric Levi-Civita symbol. The states which correspond to the same total number of quanta are degenerate. The Lie algebra of the two-dimensional x − y sub-space is generated by the diagonal operator τˆz

=

1 ( a†1 a1 − a†2 a2 ) 2

(1689)

and the ladder operators τˆ+

= a†1 a2

τˆ−

= a†2 a1

(1690)

These bi-linear τˆ operators do not change the total number of quanta and transform an energy eigenstate into eigenstates that are degenerate with it. The set of τˆ operators are linearly independent and obey similar commutation relations to the angular momentum operators. Hence, the degeneracy of the twodimensional harmonic oscillator is describable by a SU (2) symmetry group and is a Lie algebra of rank unity25 . The symmetry group of the three-dimensional harmonic oscillator must contain the SU (2) group as a sub-group. For the three-dimensional harmonic oscillator, the Lie algebra is supplemented by the diagonal operator ˆ = λ

1 √ ( a†1 a1 + a†2 a2 − 2 a†3 a3 ) 2 3

(1691)

and, therefore, is a Lie algebra of rank 2. The Lie algebra of the threedimensional harmonic oscillator also involves the following set of ladder operators βˆ+ βˆ−

= a†1 a3

γˆ+

= a†3 a2

γˆ−

= a†3 a1

= a†2 a3 (1692)

Hence, the Lie algebra contains eight generators and corresponds to the SU (3) Lie group26 . 25 The rank of a Lie algebra is given by the maximal number of (linearly independent) simultaneously commuting operators in the algebra, excluding the Casimir operators. The Casimir operators are non-linear combinations of the generators that commute with every generator of the Lie algebra. For semi-simple Lie groups of rank r, there are r Casimir operators, and their eigenvalues uniquely characterize the multiplets of the group. 26 The special unitary group in N dimensions SU (N ) contains N 2 −1 generators. The SU (N ) symmetry groups are discussed by S. Sternberg in Group Theory and Physics, Cambridge University Press, (1994).

366

4.6.23

The Bound States of the Coulomb Potential

The time-independent Schr¨odinger equation27 for the attractive Coulomb potential is     ¯h2 ∂ ¯h2 l ( l + 1 ) Ze2 2 ∂ − r + − RE,l (r) = E RE,l (r) 2 m r2 ∂r ∂r 2 m r2 r (1693) where Z is the atomic number.

0.25

V(r)/V(r0) = -r0/r

0

E3

E2 E1 -0.25

E0

-0.5

-0.75 -30

-20

-10

0

10

20

30

r/r0

Figure 93: The radially symmetric potential, V (r), for the hydrogen atom. The energy eigenvalues are denoted by horizontal lines. We introduce a dimensionless parameter r −2mE r ρ = ¯h2

(1694)

and a dimensionless constant ρ0 such that V (r) ρ0 = E ρ so that

r ρ0 =

27 E.

2 m Z e2 −E ¯ h

Schr¨ odinger, Ann. Phys. (Leipzig) 79, 361 (1926).

367

(1695)

(1696)

Then, the radial wave function is expressed in the form RE,l (r) = ρl exp[ − ρ ] f (ρ)

(1697)

which explicitly displays the limiting behavior needed at the boundaries. The factor proportional to ρl is introduced so that the wave function satisfies the Schr¨ odinger equation near the origin, where the centrifugal barrier dominates the potential. The exponential term describes the localization of the bound particle to the region around the origin. It should be noted that there is a second independent solution for which R would diverge at large ρ as exp[ + ρ ] but this second solution is to be discarded as it is not normalizable. Substituting the expression (1697) in the Schr¨odinger equation yields     ∂2f l + 1 ∂f 2 ( l + 1 ) − ρ0 + 2 − 1 − f = 0 (1698) ∂ρ2 ρ ∂ρ ρ —————————————————————————————————— Mathematical details In terms of the dimensionless variables, the equation for the radial wave function R(r) takes the form   ∂2R 2 ∂R l(l + 1) ρ0 − − + + 1 R − R = 0 (1699) ∂ρ2 ρ ∂ρ ρ2 ρ We shall substitute the radial wave function in the form R(r) = ρl exp[ − ρ ] f (ρ)

(1700)

into the above equation. The detailed form of R was chosen to satisfy our boundary conditions. The first derivative of R is calculated as   ∂R l ∂ = ρl exp[ − ρ ] − 1 + f (ρ) (1701) ∂ρ ρ ∂ρ The second derivative of R is calculated as   ∂2R l(l − 1) 2l l = ρ exp[ − ρ ] − + 1 f (ρ) ∂ρ2 ρ2 ρ   l ∂f + ρl exp[ − ρ ] 2 − 1 ρ ∂ρ 2 ∂ f + ρl exp[ − ρ ] ∂ρ2 On substituting these expressions in eqn(1699), one finds     ∂2f l + 1 ∂f 2 ( l + 1 ) − ρ0 + 2 − 1 − f = 0 ∂ρ2 ρ ∂ρ ρ 368

(1702)

(1703)

—————————————————————————————————— On multiplying by ρ, this equation reduces to the differential equation for the associated Laguerre polynomials. On multiplying by ρ, this equation reduces to the equation for the associated Laguerre polynomials L(z) in the variable z = 2 ρ,     ∂L ρ0 ∂2L + 2(l + 1) − z + − l − 1 L = 0 (1704) z ∂z 2 ∂z 2 The solution has to satisfy the boundary condition at the origin ρ = 0 and at ρ → ∞. The solution for f (ρ) can be obtained as a Fr¨obenius series X f (ρ) = an ρn (1705) n=0

On introducing this power series ansatz for the solution into the differential equation (1698) and combining terms of similar order, one finds a polynomial equation. The coefficients of the terms of various powers of ρ in the resulting polynomial must vanish identically. This procedure leads to a recursion relation which relates the coefficients an with different n. The recursion relation is found as   an+1 ( n + 1 ) ( n + 2 l + 2 ) = an 2 ( n + l + 1 ) − ρ0 (1706) If the series did not terminate, then one can show that the sum of the infinite series would be of order exp[ + 2 ρ ] for large ρ. This would result in the appearance of the second solution with the unphysical asymptotic behavior at large r   lim

r → ∞

R(r) ∼ ρl exp

+ ρ

(1707)

but this would not satisfy our boundary conditions. Therefore, one insists that the series must truncate if it is to represent a bound state. The series truncates at an integer value of n, say nr , if ρ0 = 2 ( n r + l + 1 )

(1708)

and represents a bound state with radial quantum number nr . On using r 2 m Z e2 (1709) ρ0 = − E ¯h one finds the non-relativistic formula for the energy eigenvalues E = −

Z 2 e4 m 2 ¯h ( nr + l + 1 )2 2

369

(1710)

Furthermore, on defining a characteristic length scale r0 in terms of ρ0 and eqn(1694), one has s ¯h2 r0 = ρ0 (1711) −2mE Therefore, one has r0

=

2 ( nr + l + 1 )2

¯2 h Z m e2

(1712)

and one identifies the characteristic length scale with twice the Bohr radius. The principal quantum number n is usually defined as n =

ρ0 = ( nr + l + 1 ) 2

(1713)

where l is the quantum number corresponding to the orbital angular momentum and nr is the radial quantum number. The lowest energy state or ground state corresponds to l = 0 and nr = 0. The states with higher values of E are degenerate since for these energies, different values of nr and l can be combined to give the same value of ρ0 or n. The possible values of l for a fixed ρ0 are given by l = 0, 1, 2, ..... , ρ20 − 1. Since each l value has a degeneracy of 2 l + 1 due to the different m values, the total degeneracy is ρ0 2

−1 X

 (2l + 1) =

l=0

ρ0 2

2

= n2

(1714)

Thus, the energy eigenfunctions corresponding to the bound state energy En are n2 -fold degenerate, if we neglect spin. Note that if we had deformed the Coulomb potential slightly, the degeneracy between the states with principal quantum number n but with different orbital angular momentum l would be lifted. This occurs in H, if one considers the effect of the finite size of the nucleus. The “accidental degeneracy” of the energy eigenvalues En of the hydrogen atom, corresponding to the states with principal quantum number n but with different angular momentum quantum numbers l, involves the group SO(4). More specifically, it can be shown that the energy eigenstates of the hydrogen atom can be transformed into the eigenstates of a free rotor in a four-dimensional space28 . —————————————————————————————————— 28 V.A.

Fock, Z. Phys. 98, 145 (1935).

370

0.8 0.7

l=0

n=1

0.6

Rn,0(r)

0.5 0.4 0.3 0.2

n=3

0.1 0 0

2

4

-0.1

6

8

10

12

14

16

18

20

n=2

-0.2

r/a

Figure 94: The radial wave functions for the hydrogen atom Rn,l (r), for l = 0 , 1 , 2. The radial quantum number nr determines the number of nodes in the wave function. 0.2

l=1

0.15

Rn,1(r)

0.1

n=2

0.05

n=3 0 0

2

4

6

8

10

12

14

16

18

20

-0.05

r/a

0.06

0.05

l=2

Rn,2(r)

0.04

0.03

n=3 0.02

0.01

0 0

2

4

6

8

10

371r / a

12

14

16

18

20

0.6

n=1 l=0

2

2

r R1,0 (r)

0.4

0.2

0 0

1

2

3

4

r/a0

Figure 95: The probability density P (r) = r2 | Rn,l (r) |2 for finding an electron at a distance r from the nucleus, for n = 1 , 2 , 3 and different l values.

0.25

n=2 l=1

l=0

0.15

2

2

r R2,l (r)

0.2

0.1

0.05

0 0

2

4

6

8

10

12

r/a0

0.12

l=2

n=3

l=1 l=0

2

r R3,l (r)

0.09

2

0.06

0.03

0 0

4

8

12

r/a0 372

16

20

0.5

l=1

n=2

l=0

0.4

P(r)

0.3

0.2

0.1

0 0

1

2

3

4

5

6

7

8

9

10

r / a0

Figure 96: The probability density P (r) = r2 | Rn,l (r) |2 for finding an electron at a distance r from the nucleus, for n = 2 , 3 and different l values. The classical probability densities are also shown. The classical probability densities are calculated with a fixed energy E and angular momentum l, but unknown initial position r. Only the classical particle with l = 0 reaches the origin, however, it has a vanishing probability density there as the speed tends to infinity as r → 0.

0.2

n=3 l=2 l=1

0.15

P(r)

l=0 0.1

0.05

0 0

2

4

6

8

10

r / a0

373

12

14

16

18

20

4.6.24

Exercise 95

Derive the recursion relation for the series expansion of the associated Laguerre polynomials that occur in the radial wave functions of electrons in the bound states of Hydrogen atoms and determine the condition that the series terminates after n terms. ——————————————————————————————————

4.6.25

Exercise 96

An atom of tritium is in its ground state when the nucleus suddenly decays into a helium nucleus along with the emission of a fast electron (and an antineutrino) which leaves the atom without perturbing the extra-nuclear electron. Find the probability that the remaining He+ atom will be left in the state with n = 1, l = 0 and n = 2, l = 0. What is the selection rule for the l quantum number? ——————————————————————————————————

4.6.26

Solution 96

The electronic state immediately before the decay is identical to that of the n = 1, l = 0 state of H, i.e. Z = 1. Then,   1 − 3 r ΨZ=1,n=1,l=0 (r) = √ a0 2 exp − (1715) a0 π The probability that the electron falls into the energy levels of He+ is given by the overlap with the respective Z = 2 wave functions 1 ΨZ=2,n=1,l=0 (r) = √ π



2 a0

 32

 exp

− 2

r a0

 (1716)

and 1 ΨZ=2,n=2,l=0 (r) = √ π



1 a0

 32 

r 1 − a0



 exp

r − a0

 (1717)

The matrix elements between the Z = 1 and Z = 2 states are evaluated via scaling and repeated integration by parts as Z ∞ Cn=1,l=0 = 4 π dr r2 Ψ∗Z=1,n=1,l=0 (r) ΨZ=2,n=1,l=0 (r) 0 1

=

82 4π π a30

Z

∞ 2

dr r exp 0

374



r − 3 a0



22

=

22 33



Z

7

=

dx x2 exp



 − 3x

0 9

(1718)

Thus, the probability of the electron falling into the ground state is equal to P (0) =

29 36

(1719)

Similarly, ∞

Z

Cn=2,l=0

dr r2 Ψ∗Z=1,n=1,l=0 (r) ΨZ=2,n=2,l=0 (r)     Z ∞ r r 1 2 = 4π dr r 1 − exp − 2 π a3 0 a0 a0     Z ∞0 = 4 dx x2 1 − x exp − 2 x =



0

0

1 = − 2

(1720)

Thus, the probability of the electron falling into the first excited state is equal to 1 (1721) P (1) = 2 2 The selection rule is ∆l = 0. This follows as the potential is spherically symmetric, and so angular momentum is conserved. Mathematically this is manifested by the angular part of the wave function being given by spherical harmonics, and also by the fact that the spherical harmonics with different l are orthogonal. ——————————————————————————————————

4.6.27

Exercise 97

A tritium atom decays via the nuclear reaction 3

H →

3

He + e− + ν

(1722)

and a fast electron leaves the atom almost instantaneously. Calculate the expectation value of the energy for the remaining electron. The probability that the remaining energy is in the n-th bound state is denoted by P (n). It is found that ∞ X P (n) = 0.9755 (1723) n=1

and

∞ X n=1

P (n)

1 = 0.76660 n2 375

(1724)

What is the probability that the Helium ion is doubly ionized and what is the average energy of the second emitted electron? —————————————————————————————————— 4.6.28

Ladder Operators for the Hydrogen Atom

The radial equation for the energy eigenstates of Hydrogen atom can be written in a dimensionless form. This is achieved by introducing a length scale b defined by ¯h2 (1725) b = Z e2 m and a dimensionless variable ρ defined by ρ =

r b

The energy eigenvalue can then be expressed in dimensionless form as   2 m b2  = E ¯h2   2 ¯h2 = E Z 2 e4 m   E = R

(1726)

(1727)

Thus, the dimensionless energy  corresponds to the value of E when expressed in units of Rydbergs. In dimensionless units, the radial equation takes the form   l(l + 1) 2 ∂2 + − un,l =  un,l (1728) − ∂ρ2 ρ2 ρ where un,l (ρ) = ρ Rn,l (ρ)

(1729)

is the reduced radial wave function. The bound state wave functions have negative eigenvalues, are normalizable, satisfy the boundary condition limρ→0 un,l (ρ) → ρl+1 at ρ = 0, and vanish asymptotically as   √ lim un,l (ρ) → exp − − ρ (1730) ρ→∞

when ρ → ∞. The radial equation can be solved for the bound states using appropriate raising and lowering operators. The lowering operator Aˆ− l is defined as ∂ l + 1 1 Aˆ− + − l = ∂ρ ρ l + 1 376

(1731)

and the raising operator Aˆ+ l is defined as the Hermitean conjugate of the lowering operator l + 1 1 ∂ Aˆ+ + − (1732) l = − ∂ρ ρ l + 1 It immediately follows that   ∂2 l(l + 1) 2 1 ˆ+ = Aˆ− A − + − + (1733) l l ∂ρ2 ρ2 ρ ( l + 1 )2 so the radial Schr¨ odinger equation can be written as   1 − ˆ+ ˆ Al Al − un,l =  un,l ( l + 1 )2

(1734)

The product of the raising and lowering operators, taken in the reverse order, can be expressed as   (l + 2)(l + 1) 2 1 ∂2 + ˆ− ˆ + − + (1735) Al Al = − ∂ρ2 ρ2 ρ ( l + 1 )2 Hence, one has the relation ˆ− Aˆ+ l Al −

1 1 ˆ+ = Aˆ− l+1 Al+1 − 2 (l + 1) ( l + 2 )2

(1736)

On multiplying the radial Schr¨odinger equation (as expressed in the form of eqn(1734)) by the raising operator Aˆ+ l , one has   1 + − ˆ+ ˆ ˆ Al Al Al − un,l =  Aˆ+ l un,l ( l + 1 )2   1 ˆ− Aˆ+ =  Aˆ+ (1737) Aˆ+ l un,l l un,l l Al − ( l + 1 )2 On substitution of the relation expressed by eqn(1736) in the above equation, one finds   1 ˆ+ − Aˆ− A Aˆ+ =  Aˆ+ (1738) l+1 l+1 l un,l l un,l ( l + 2 )2 Hence, we see that Aˆ+ odinger equation l un,l is an eigenstate of the radial Schr¨ with angular momentum l0 = l + 1 and energy eigenvalue . Thus, Aˆ+ l is a raising operator for the radial wave function. The eigenstates with angular momentum l0 = l − 1 can be obtained with the aid of the lowering operator. This can be shown by considering the Schr¨odinger equation written in the form   1 + − ˆ ˆ Al−1 Al−1 − 2 un,l =  un,l (1739) l 377

On applying Aˆ− l−1 to this equation, one finds   ˆ+ Aˆ− − 1 Aˆ− A un,l l−1 l−1 l−1 l2   1 ˆ+ Aˆ− Aˆ− l−1 Al−1 − 2 l−1 un,l l   1 ˆ− Aˆ+ Aˆ− l−2 Al−2 − l−1 un,l ( l − 1 )2

=  Aˆ− l−1 un,l =  Aˆ− l−1 un,l =  Aˆ− l−1 un,l

(1740)

where we have used eqn(1736) in the second line. On comparison of eqn(1740) with eqn(1739), one finds that Aˆ− l−1 un,l is an eigenstate with eigenvalue  and angular momentum l0 = l − 1. The energy eigenvalue  satisfies an inequality which can be found directly from the Radial Schr¨ odinger equation   1 − ˆ+ ˆ Al Al un,l = +  un,l (1741) ( l + 1 )2 by multiplying by u∗n,l and integrating over ρ 29   Z ∞ Z ∞ 1 ∗ ˆ+ un,l = dρ u∗n,l Aˆ− A dρ u +  un,l n,l l l ( l + 1 )2 0 0

(1742)

ˆ+ However, un,l is normalized to unity and Aˆ− l is the Hermitean conjugate of Al , so   Z ∞ 1 2 dρ | Aˆ+ u | = +  (1743) n,l l ( l + 1 )2 0 Since the left-hand side is positive, one finds the inequality  ≥ −

1 ( l + 1 )2

(1744)

 ≥ −

1 ( l + 1 )2

(1745)

The inequality

involving the (negative) bound state energy eigenvalue  and the angular momentum indicates that the angular momentum raising process must terminate. If this terminates for some value of l, say lmax , then the termination condition is expressed as Aˆ+ (1746) lmax un,lmax = 0 29 Due to the factor of ρ which appears in the definition of u n,l in terms of the radial function Rn,l (ρ), the three-dimensional radial equation is reduced to an effective one-dimensional Schr¨ odinger equation. The factor of ρ in un,l also reduces the appropriate measure for the radial integration is reduced from ρ2 to unity, as expected for an effective one-dimensional geometry.

378

On using the termination condition in the eigenvalue equation   1 ˆ+ Aˆ− A − un,lmax =  un,lmax lmax lmax ( lmax + 1 )2

(1747)

so that the first term in the parenthesis is zero, one finds  = − or

1 ( lmax + 1 )2

(1748)

1 n2 + 1 ) is the principal quantum number.  = −

where n = ( lmax

(1749)

The radial functions un,l can be obtained by first solving the first order differential equation for un,lmax that expresses the termination and then, using the lowering operator successively to find all the eigenfunctions with lower l values. The termination condition Aˆ+ lmax un,lmax = 0

(1750)

is expressed as ∂un,lmax = ∂ρ



n 1 − ρ n

 un,lmax

(1751)

ρ n

(1752)

This can be integrated to yield ln un,lmax = A + n ln ρ −

where A is a constant of integration. On exponentiating, one finds the radial wave function has the form   ρ un,lmax = C ρn exp − (1753) n where the constant C has to be determined from the normalization condition. The above form of un,l (ρ) yields a solution for the radial wave function which does satisfy the correct boundary conditions. The radial wave function with angular momentum l = lmax − 1 is obtained from un,lmax by the action of the lowering operator un,lmax −1 ∝ Aˆ− lmax −1 un,lmax

(1754)

or, more explicitly 

un,lmax −1

 ∂ (n − 1) 1 ∝ + − un,lmax ∂ρ ρ (n − 1)     ρ ρ n−1 ∝ 1 − ρ exp − n(n − 1) n 379

(1755)

Table 8: Radial wave functions Rn,l (ρ) for the Coulomb where ρ = R ∞potential, Z r 2 . The Radial wave functions are normalized so that dρ ρ R (ρ)2 = 1. n,l a 0  n = 1

n = 2

l = 0

√1 2

l = 0

l = 0



2 3

32



7

32

1 √

2 3

1 −

22

6



1 −

22 9 √

ρ 6

2

5

2 27

ρ 2



ρ

2

ρ 2



exp

ρ exp

ρ +



32





3

l = 2

ρ 2

1 −

2

5

l = 1

− ρ

2 exp

l = 1

n = 3









 ρ exp

ρ exp



exp





 −

− ρ 3

ρ 3

ρ 3







Therefore, the radial functions corresponding to smaller values of l can be obtained by sequentially operating with the appropriate Aˆ− l−1 .

4.6.29

Rydberg Wave Packets

The correspondence between the classical picture of the orbital motion of electrons in the hydrogen atom and the quantum mechanical picture can be seen by examining the time dependence of the peak in the probability density | Ψ(r, θ, ϕ; t) |2 . For convenience, we shall consider motion in the plane θ = π2 , and so shall set m to the maximal value m = l. In this case, due to the θ dependence given by Θll (θ) ∝ sinl θ, the probability density is maximal at θ = π2 . To obtain a time dependence of the probability, it is necessary to superimpose energy eigenstates to produce a wave packet. In particular, to obtain a probability density that shows a time-dependent variation in the ϕ dependence, eigenstates corresponding to different energy eigenvalues must be superimposed. We shall superimpose the energy eigenstates with l = n − 1, which are given by  n−1     r r Ψn,n−1,n−1 (r, θ, ϕ) = ℵ−1 exp − sinn−1 θ exp i ( n − 1 ) ϕ a0 n a0 (1756) 380

2

where a0 = mh¯ e2 is the Bohr radius, and ℵ is the normalization. This set of energy eigenstates correspond to circular orbits, as the radial quantum number nr = 0. For the circular orbits, the principal quantum number n is just given by l + 1. The energy eigenvalues are given by En = −

m e4 2 ¯h2 n2

These states have radii defined by the maxima in 2 r Ψn,n−1,n−1 (r, θ, ϕ)

(1757)

2

(1758)

which occur at r = n2 a0 . The products of the uncertainties in the radial position and the radial momentum are given by   1 ¯h ( ∆r )rms ( ∆pr )rms = 1 + + ... (1759) 2 2n and also for the angular coordinates ( ∆θ )rms ( ∆pθ )rms

¯ h = 2



1 1 + + ... 4n



A particular wave packet is represented by the wave function   X it Ψ(r; t) = Cn Ψn,n−1,n−1 (r, θ, ϕ) exp − En ¯h n

(1760)

(1761)

where the distribution of Cn produces a mean value of n denoted by n, and a width given by ∆nrms . The probability density shows that the wave packet describes periodic circular orbits, with period given by the Kepler period TK . This is seen in fig(98). This can be derived from the expansion of the exponential factor in the wave packet about n        t t ∂En exp i ( n ϕ − En ) exp i ( n − n ) ϕ − . . . (1762) h ¯ ¯h ∂n n=n The first factor being a common overall phase factor does not appear in the probability density. If the terms of order ( n − n )2 can be neglected, the sum over n of the second factor produces a finite Fourier series. The resulting approximation to the wave packet is 2 π periodic in the variable   t ∂En (1763) ϕ − ¯h ∂n n=n

381

0.05

n=6 l=5

0.04

l=0

P(r)

0.03

0.02

0.01

0 0

10

20

30

40

50

60

70

80

90

100

r / a0

Figure 97: A comparison of the quantum mechanical probability densities of the circular orbit with l = n − 1 (blue) and the linear orbit l = 0 (red). For the circular orbit with n = l + 1, the zero-point radial motion becomes negligible for large values of l and the radial wave function becomes highly peaked at the radius of the orbit. The classical probability densities are denoted by the broken lines. Thus, the Kepler period is given by 2π TK



=

1 ¯h

∂En ∂n

=

m e4 ¯h3 n3



n=n

(1764)

or TK

=

2 π ¯h3 n3 me4

h ¯ 2 n2 m e2

= 2π

=



e2 h ¯ n

2 π a0 n2 v





(1765)

which is the size of the circular orbit divided by the average speed v. The 382

average speed is given in terms of the fine structure constant and the velocity of light c via v

= =

e2 ¯h n c e2 n ¯h c

(1766)

The terms of order ( n − n )2 have the effect of dephasing the oscillations. At first, the ϕ dependence of the wave packets distort by spreading slowly in time. This increases the range of ϕ in which the wave packets have an appreciable magnitude. The initial spreading is similar to the spreading of a wave packet for a free particle. However, at large times, the wave packet recurs or re-forms as is shown in fig(99). The time scale for this phenomenon is found by expanding the energy about the average value of n, 2 ∂E 1 2 ∂ E ( n − n ) (1767) E n = En + ( n − n ) + 2 ∂n2 n=n ∂n n=n The recurrence time TR is defined via −

1 ∂ 2 E 2π = TR ¯h ∂n2 n=n

(1768)

which yields n TK (1769) 3 The recurrence time causes the wave function to dephase and recur at time t = TR but with a phase shift of π. The phenomenon of recurrence is a quantum effect involving the non-linear dependence of the energy eigenvalues30 . TR =

4.6.30

Laguerre Polynomials

Laguerre’s differential equation is ∂ 2 φn ∂φn + (1 − z) + n φn = 0 2 ∂z ∂z The solution can be represented by a contour integral   I dt zt 1 φn (z) = exp − 2πi 1 − t ( 1 − t ) tn+1 z

(1770)

(1771)

where the contour runs around a circle centered on the origin, with radius less than unity. We shall show that this is a solution of the differential equation. First we shall evaluate the derivatives   I ∂φn (z) dt zt t exp − (1772) = − ∂z 2πi 1 − t ( 1 − t )2 tn+1 30 C. R. Stroud Jr. Physics and Probability, Essays in honor of E. T. Jaynes, Grandy and Milonini editors, Cambridge University Press (1993).

383

Figure 98: The time dependence of the probability density for an electron in a circular orbit, at times of order TK . The wave packet rotates with period Tk . [After C. R. Stroud Jr. (1993).] and ∂ 2 φn (z) = + ∂z 2

I

dt exp 2πi

 − 384

zt 1 − t



t2 ( 1 − t )3 tn+1

(1773)

Figure 99: The time dependence of the probability density for an electron in a circular orbit, at large times. At large times the wave packet recurs. [After C. R. Stroud Jr. (1993).] Then, on substituting the derivatives into the differential equation, one obtains     I dt zt 1 2 2 0 = exp − zt −t(1−t)(1−z)+n(1−t) 2πi 1 − t ( 1 − t )3 tn+1 (1774) The right hand side is identified as the integral of a perfect differential     I zt 1 dt ∂ exp − (1775) 0 = − 2 π i ∂t 1 − t ( 1 − t ) tn On integrating the perfect differential around a closed contour without crossing a branch point, we find zero. Thus, we have verified that the expression given by the contour integral does satisfy Laguerre’s equation. We shall denote the solution of Laguerre’s equation by   I dt zt 1 Ln (z) = exp − 2πi 1 − t ( 1 − t ) tn+1

(1776)

The generating function for the Laguerre polynomials, G(z, t) can be found directly from this representation. Let G(z, t) =

∞ X n=0

385

Ln (z) tn

(1777)

then divide by tn+1 and integrate around a closed contour containing the origin. This procedure yields

Contour of Integration Im t

2

1

0 -2

-1

0

1

2

Re t -1

-2

Figure 100: The contour for the integration in eqn(1776). The circular contour encloses the pole at the origin but excludes the pole at t = 1. The two poles of the integrand are marked by the crosses. I

dt G(z, t) 2 π i tn+1

=

X

I Lm (z)

m

dt tm 2 π i tn+1

= Ln (z)

(1778)

since Cauchy’s theorem implies only the term with n = m is non-zero. On comparing the first line with the integral expression for Ln (z), we identify   z t exp − 1 − t G(z, t) = (1779) (1 − t) Thus, we have the generating function series expansion   exp − 1 z−t t ∞ X = Ln (z) tn (1 − t) n=0 for t < 1.

386

(1780)

A Rodriguez formula for the solution can be found by performing the nonlinear transformation tz = s − z (1781) 1 − t which can be solved for t as s − z t = (1782) s In this variable, the contour integral in eqn(1776) takes on the form I ds sn exp[ − s ] Ln (z) = exp [ z ] (1783) 2 π i ( s − z )n+1 where the new contour encloses the point s = z which corresponds to t = 0. By Cauchy’s theorem, this integral can be seen to be the n-th order derivative of the numerator evaluated at the pole   exp [ z ] ∂ n n Ln (z) = z exp[ − z ] (1784) n! ∂z n for integral n. The Laguerre differential operator is not Hermitean, but if the solution is expressed as   z Φn (z) = exp − Ln (z) (1785) 2 then the corresponding operator is Hermitean, and the differential equation becomes ∂Φn 1 z ∂ 2 Φn + + (n + − ) Φn (z) = 0 (1786) z ∂z 2 ∂z 2 4 Therefore, the Laguerre polynomials satisfy the orthogonality condition   Z ∞ dz exp − z Lm (z) Ln (z) = δm−n (1787) 0

The Associated Laguerre polynomials are defined by Lkn (z) = ( − 1 )k

∂k Ln+k (z) ∂z k

(1788)

For fixed k, one can find a generating function for Lkn (z) from the generating function for the Laguerre polynomials   exp − 1 z−t t ∞ X Ln+k (z) tn+k (1789) = (1 − t) n=−k

387

by differentiating k times, with respect to z, obtaining   z t exp − 1 − t tk ∞ X ∂k = ( − 1 )k Ln+k (z) tn+k k+1 (1 − t) ∂z k

(1790)

n=−k

Hence, since Ln+k (z) is a polynomial with highest order term of z n+k , the sum has its first non-zero term when n = 0 and   z t exp − 1 − t ∞ X = Lkn (z) tn (1791) ( 1 − t )k+1 n=0 which is the generating function expansion for the associated Laguerre polynomials. The expansion also shows that the associated Laguerre polynomials reduce to the Laguerre polynomials when k = 0. The associated Laguerre polynomials satisfy the differential equation z

∂ k ∂2 k Ln + ( k + 1 − z ) L + n Lkn = 0 2 ∂z ∂z n

(1792)

The associated Laguerre polynomials have the Rodriguez representation       n 1 k − k ∂ n + k Ln (z) = exp + z z exp − z z (1793) n! ∂z n The differential equation for the associated Laguerre polynomials can be written in the form of a Hermitean operator eigenvalue equation by using the substitution   k z φkn (z) = z 2 exp − Lkn (z) (1794) 2 The functions φkn (z) form a complete orthonormal set because they satisfy the differential equation   ∂ 2 φkn ∂φkn z 2n + k + 1 k2 z + − − + φkn = 0 (1795) ∂z 2 ∂z 4 2 4z The normalization integral is   Z ∞ ( n + k )! k dz z exp − z Lkn (z) Lkm (z) = δn−m n! 0

The normalized hydrogen atom wave function is given by   l l Ψn,m,l (r, θ, ϕ) = Cl,n,m ρ exp − ρ L2l+1 n−l−1 ( 2 ρ ) Ym (θ, ϕ) 388

(1796)

(1797)

Table 9: The Associated Laguerre Polynomials Lkn (z) k=0

k=1

n

L0n (z)

L1n (z)

n=0

1

1

n=1

1−z

2−z

1 2 (2

n=2 n=3

1 6 (6

1 2 (6

− 4z + z 2 )

− 18z + 9z 2 − z 3 )

1 6 (24

− 6z + z 2 )

− 36z + 12z 2 − z 3 )

k=2

k=3

n

L2n (z)

L3n (z)

n=0

1

1

n=1

3−z

4−z

n=2 n=3

1 2 (12 1 6 (60

1 2 (20

− 8z + z 2 )

− 60z + 15z 2 − z 3 )

389

1 6 (120

− 10z + z 2 )

− 90z + 18z 2 − z 3 )

where we have written ρ =

Z r n a0

(1798)

and a0 is the Bohr radius. The normalization constant is given by  3 ( n − l − 1 )! 2Z | Cl,n,m |2 = 22l n a0 2 n ( n + l )!

(1799)

——————————————————————————————————

4.6.31

Exercise 98

Show that for the hydrogen atom, the average values of the moments of the radial distance of the electron from the proton satisfy the following relations,   4(s + 1) s s 2 2 s−1 + r − ( 2 s + 1 ) a r ( 2 l + 1 ) − s a20 rs−2 = 0 0 ρ20 4 s > −(2l + 1) (1800) where a0 is the Bohr radius. Also show that r r−1

ρ20 4 4 a0 ρ20

= a0 =

 1 +

1 2

 1 − 4

l(l + 1) ρ20



(1801)

——————————————————————————————————

4.6.32

Solution 98

We shall use the dimensionless variable ρ where r −2mE r ρ = κr = ¯h2 and write the Radial part of the wave function as R(ρ) = introduce a constant ρ0 by Z e2 ρ0 − = Er ρ 390

(1802) f (ρ) ρ .

We also (1803)

Then, the radial part of the energy eigenvalue equation becomes  2    l(l + 1) ρ0 ∂ f + − 1 f = 0 + − ∂ρ2 ρ2 ρ

(1804)

We shall multiply this equation by ρs+1

∂f − C ρs f ∂ρ

(1805)

and integrate over ρ. Then, as the average of the powers of ρ are given by Z ∞ s ρ = dρ f (ρ) ρs f (ρ) (1806) 0

this will give us terms that are equal to the required expectation values. The terms proportional to C are evaluated as   C l ( l + 1 ) ρs−2 − ρ0 ρs−1 + ρs Z ∞ ∂2f −C dρ ρs f ∂ρ2 0

(1807)

The last term can be integrated by parts, and the boundary terms vanish for appropriately large l. Then we obtain   C l ( l + 1 ) ρs−2 − ρ0 ρs−1 + ρs Z +C

∞ s

dρ ρ f 0



∂f ∂ρ

2 − C

s ( s − 1 ) s−2 ρ 2

The remaining terms are re-written as  2   Z ∞ ∂f ∂ f l(l + 1) ρ0 dρ ρs+1 + − + − 1 f ∂ρ ∂ρ2 ρ2 ρ 0    Z ∞ 1 l(l + 1) ρ0 ∂f 2 s+1 = dρ ρ − + − 1 2 0 ρ2 ρ ∂ρ   2  Z ∞ ∂f ∂ f + dρ ρs+1 ∂ρ ∂ρ2 0 These terms are evaluated as   1 l ( l + 1 ) ( s − 1 ) ρs−2 − s ρ0 ρs−1 + ( s + 1 ) ρs 2  2 Z ∞ (s + 1) ∂f s − dρ ρ 2 ∂ρ 0 391

(1808)

(1809)

(1810)

 The terms involving

∂f ∂ρ

2 vanish if we choose the constant C as

C =

(s + 1) 2

(1811)

As the sum of the two terms are equal to zero, due to F satisfying the differential equation, one has the equation relating the expectation values of powers of ρ   1 2s + 1 s l ( l + 1 ) − ( s2 − 1 ) ρs−2 − ρ0 ρs−1 + ( s + 1 ) ρs = 0 4 2 (1812) On putting s = 0 and noting ρ0 = 1 we have ρ−1 =

2 ρ0

(1813)

and then with s = 1 we have   3 1 2 2ρ = ρ0 − ( 2 l + 1 ) − 1 ρ−1 2 4   1 3 2 ρ0 − ( 2 l + 1 ) − 1 ρ−1 = 0 2 2

(1814)

On converting back from dimensionless variables, we find the result given. ——————————————————————————————————

4.6.33

Exercise 99

Find the momentum space wave functions for the two lowest m = 0 energy eigenstates of the hydrogen atom, given by 1 Ψ1,0,0 (r) = √ π



Z a0

 32

 −

exp

Z r a0

 (1815)

and r Ψ2,1,0 (r) =

3 4 π 4!



Z a0

 32 

Z r a0



 exp



Z r 2 a0

 cos θ

(1816)

——————————————————————————————————

392

4.6.34

Solution 99

The momentum space wave function is given by the Fourier transform   32 Z   1 i 3 Φ(p) = d r exp − p . r Ψ(r) (1817) 2π¯ h ¯h For the (1, 0, 0) wave function, the angular integration can be easily performed leading to   32 √ Z Φ1,0,0 (p) = 2 π ¯h 2 π ¯h a0      exp + h¯i p r − exp − h¯i p r    Z ∞ Z r 2 exp − × dr r ipr a0 0 (1818) The first factor in the integrand is recognized as being proportional to the l = 0 spherical Bessel function. The integral is easily evaluated to yield "   32 2  2 # √ i 1 Z ¯ h 1 Φ1,0,0 (p) = 2 π − 2 π a0 p p + i Za0h¯ p − i Za0h¯ # "  32  Z h ¯ √ Z ¯ h a0 (1819) = 8 π  2 2 π a0 Z h ¯ 2 2 p + ( a0 )

The second momentum space wave function is found by using the Rayleigh expansion of the exponential plane wave. The Legendre polynomial is given in terms of the angle between the momentum and the position vector. The spherical harmonic addition theorem is used to expand the Legendre polynomials in terms of the products of the spherical harmonics, each separately involving the directions of the momentum and position. On integrating over the directions of the position vector and using orthonormality, one finds that the momentum space wave function depends upon the spherical harmonic with the same (l, m) as the real space wave function, except that it depends on the direction of the momentum, (θp , ϕp ), with respect to the Cartesian axes. The coefficient of the spherical harmonic Yml (θp , ϕp ) is given by an integration involving the product of the spherical Bessel function jl (kr) and the radial dependence of the real space wave function. The result entails evaluating integrals of the form   Z ∞ (2+m) dz z exp − α z jl (z) (1820) 0

which for l = 2 simplifies to    Z ∞ sin z − z cos z (2+m) dz z exp − α z z2 0 393

Table 10: The Gegenbauer Functions Cνµ (x).

ν ν ν ν ν ν

=0 =1 =2 =3 =4 =5

µ=1

µ=2

µ=3

µ=4

µ=5

1 2x 4x2 − 1 8x3 − 4x 4 16x − 12x2 + 1 32x5 − 32x2 + 6x

1 4x 12x2 − 2 32x3 − 12x 80x4 − 48x2 + 3

1 6x 24x2 − 3 80x3 − 24x

1 8x 40x2 − 4

1 10x

= ( − 1 )m = ( − 1 )m

   Z ∞ ∂m dz exp − α z sin z − z cos z ∂αm 0   2 ∂m (1821) ∂αm ( 1 + α2 )2

The momentum space wave function of a general energy eigenstate of a Hydrogen-like atom has been calculated by Podolsky and Pauling31 . The momentum eigenfunctions Φn,l,m (p, θ, ϕ) are found to be given by the general expression   32  1   n a0 n(n − l − 1)! 2 l l 2l+4 Φn,l,m (p, θ, ϕ) = Ym (θ, ϕ) − (−i) π2 l! Z 2 π ¯h (n + l)!   l 2 ξ ξ −1 × 2 C l+1 (1822) (ξ + 1)l+2 n−l−1 ξ 2 + 1 where

n p a0 (1823) Z ¯h and Cνµ (x) are the Gegenbauer functions. Some of the Gegenbauer functions are given in Table(10), and the others can be generated from the recursion relation   2µ µ+1 µ+1 µ Cν (x) = x Cν−1 (x) − Cν−2 (x) (1824) ν ξ =

The momentum space distribution of the electrons in Hydrogen was determined experimentally32 . The experiments involved the ionization of atomic hydrogen by a beam of high-energy electrons. —————————————————————————————————— 31 B. 32 B.

Podolsky and L. Pauling, Phys. Rev. 34, 109 (1929). Lohman and E. Weigold, Phys. Letts. 86 A, 139 (1981).

394

Momentum Space Probability Distribution 4

n=2,l=0

4 π Pn,l(k) k2

3

n=2,l=1 (spherically averaged)

2

n=1,l=0 1

0 0

0.5

1

1.5

2

ka

Figure 101: The spherically averaged momentum-space distribution function for the few lowest energy eigenstates of Hydrogen. 4.6.35

Exercise 100

Find the scalar potential and vector potential produced by an electron in a l = 1 state of a hydrogen atom by first calculating the charge and current densities and then by using Maxwell’s equations. ——————————————————————————————————

4.6.36

Solution 100

The hydrogen atom in a 2p orbital has a charge distribution ρ(r, t)

= | Ψ(r, t) |2   e r 2 = r exp − sin2 θ 64 π a50 a0 2

(1825)

where a0 is the Bohr radius mh¯ e2 . In this case, the charge density is independent of time. The electrostatic potential φ(r) is given by Z 1 ρ(r0 ) (1826) φ(r) = d3 r 0 | r − r0 | 4π

395

We shall first perform the angular integral. The denominator is expanded in terms of the Legendre Polynomials via the expansion  l 1 1 X r< = Pl (cos γ) (1827) | r − r0 | r> r> l

where γ is the angle between r and r0 . Then using the addition theorem one has  l m=+l X 1 X r< 4π 1 = Ym∗ l (θ0 , ϕ0 ) Yml (θ, ϕ) (1828) | r − r0 | r> r> 2l + 1 l

m=−l

The angular dependence of the numerator can be expressed in terms of the spherical harmonics sin2 θ0

= =

( 1 − cos2 θ0 ) r 2 √ 4π 2 2 0 0 4 π Y00 (θ0 , ϕ0 ) − Y (θ , ϕ ) 3 5 3 0

(1829)

The angular integration can be performed using the orthogonality of the spherical harmonics. Hence,   Z ∞   e 1 r0 0 04 φ(r) = dr r exp − 64 π a50 r> a0 0 r   X  r < l 1 2 √ 4π 2 l Y (θ, ϕ) 4 π δl,0 − δl,2 × r> 2l + 1 0 3 5 3 l     2 e 1 3 cos2 θ − 1 = I0 − I2 (1830) 3 64 π a50 5 2 where

Z



dr0 r04

I0 =



0

and Z I2 =

∞ 0

dr r 0

04



1 r>



2 r< 3 r>



r0 a0



r0 − a0



 −

exp  exp

(1831)

(1832)

The radial integrations are broken into two parts, one for r > r0 and the other for r < r0 . One finds      2  3  4    a50 r r r r r I0 = 24 − 24 + 24 + 12 + 4 + exp − r a0 a0 a0 a0 a0     2  3  4    r r r r r + 6 + 6 + 3 + exp − (1833) a0 a0 a0 a0 a0 396

and I2

a70 r3







 r = 6! − 6! + 6! + a0  3  4 r r + 5! + 30 + 6 a0 a0   5  6   r r + + exp a0 a0

 2 6! r + 2 a0  5  6    r r r + exp − a0 a0 a0  r − (1834) a0

Hence, for r  a0 where all the exponential terms are suppressed, one finds that the electrostatic potential reduces to   e 1 a2 3 cos2 θ − 1 φ(r) = − 6 30 (1835) 4π r r 2 which is interpreted as the sum of a charge monopole and a quadrupole term. The deviation of the electron wave function from spherical symmetry has set up a quadrupolar electric field. When r < a0 , the potential is reduced from that of the monopole of charge e, since a Gaussian surface enclosing the origin only contains a fraction of the total electron charge. To find the vector potential, first we find the current density j. The current density is found from the expression   e¯ h j(r, t) = Ψ∗ (r, t) ∇Ψ(r, t) − Ψ(r, t) ∇Ψ∗ (r, t) (1836) 2mi In spherical polar coordinates, the gradient is written as ∇ = eˆr

∂ 1 ∂ 1 ∂ + eˆθ + eˆϕ ∂r r ∂θ r sin θ ∂ϕ

Therefore, the current density of the p electron is evaluated as     e ¯h r j(r, t) = eˆϕ exp − r sin θ 64 π m a50 a0

(1837)

(1838)

The current density is resolved into Cartesian components, via eˆϕ0 = cos ϕ0 eˆy − sin ϕ0 eˆx

(1839)

and is then expressed in terms of spherical harmonics. The result is     e¯ h r j(r, t) = i exp − r × 64 π m a50 a0 r   8π eˆx − i eˆy eˆx + i eˆy 1 × Y11 (θ, ϕ) − Y−1 (θ, ϕ) 3 2 2 (1840) 397

which is time independent. The vector potential is given by Z j(r0 ) 1 A(r) = d3 r 0 4π | r − r0 |

(1841)

Then the Legendre polynomial generating function expansion and the spherical harmonic addition theorem are used to perform the angular integral over the variables (θ0 , ϕ0 ). The integrations utilize the orthonormality of the spherical harmonics. The end result of the integrations over the angles (θ0 , ϕ0 ) is   Z ∞     e¯ h r< r0 0 03 A(r) = i dr r exp − × 2 64 π m a50 r> a0 0 r   8π 1 eˆx − i eˆy eˆx + i eˆy 1 × Y11 (θ, ϕ) − Y−1 (θ, ϕ) 3 3 2 2   Z ∞     0 1 e ¯h r< r 0 03 = eˆϕ dr r exp − sin θ 2 3 64 π m a50 r> a0 0 (1842) The vector potential has a directional dependence which is similar to the current density. The radial integration can be performed, leading to the vector potential being dominated by a dipole contribution at large distances. ——————————————————————————————————

398

4.7

A Charged Particle in a Magnetic Field

Consider a particle of mass m and charge q moving in a uniform magnetic field B oriented along the z axis. B = eˆz Bz (1843) The vector potential A(r) can be found as a solution of B = ∇ ∧ A(r)

(1844)

A(r) = − y Bz eˆx

(1845)

One solution is given by which corresponds to a particular choice of the gauge. The Hamiltonian of the charged particle in the field can be written as  2   1 q 1 ˆ = H pˆx − Ax (r) + pˆ2y + pˆ2z 2m c 2m  2   1 q 1 2 2 = pˆx + y Bz + pˆy + pˆz (1846) 2m c 2m We note that the x and z components of the momentum commute with the ˆ . Thus, Hamiltonian, as x and z do not appear in H ˆ , pˆx ] = 0 [H ˆ , pˆy ] = 0 [H

(1847)

which means that they are constants of motion. As the energy eigenstates are also eigenstates of pˆx and pˆz , we shall write the energy eigenfunction as   1 px x + pz z φn,px ,py (r) = exp i φn (y) (1848) 2 π ¯h ¯h The energy eigenvalue equation can then be written in terms of the unknown function φn (y), as  2  2     2 pˆy m q Bz c px p2 φn (y) = E − z + y+ φn (y) (1849) 2m 2 mc q Bz 2m This equation looks like the energy eigenvalue equation for the one-dimensional harmonic oscillator, in which we have shifted the origin of the y coordinate px through a distance qc B . The frequency of the oscillator corresponds to the z Larmor precession frequency ωL , which is given by  2 q Bz 2 ωL = (1850) mc Using our previously gained knowledge of the one-dimensional harmonic oscillator, we find that the energy eigenvalues are then given by   1 p2z En,pz = h ¯ ωL n + + (1851) 2 2m 399

We note that these energy levels (known as Landau levels33 ) are degenerate, as the introduction of the B field causes states with different values of the (quasicontinuous quantum numbers) px and py to collapse onto states where the value of the energy is determined by the integer n. The wave functions with the different values of px have the “oscillations” in the y coordinate centered around different points. ——————————————————————————————————

4.7.1

Exercise 101

Show that both x0 and y0 commute with the Hamiltonian, where x ˆ0 yˆ0

c pˆy q Bz c pˆx = − q Bz

= x ˆ +

(1852)

but x0 and y0 do not commute, [x ˆ0 , yˆ0 ] 6= 0

(1853)

Show that in the classical limit, these variables correspond to the projection of the particle’s orbit on the x - y plane. Hence, x0 and y0 cannot be simultaneously known for the quantum system. ——————————————————————————————————

4.7.2

Exercise 102

Consider a particle of mass m and charge e in perpendicular uniform electric and magnetic fields, E and B. Find the eigenfunctions and eigenvalues. Find the average velocity in the x direction for any eigenstate. ——————————————————————————————————

4.7.3

Solution 102

Let the vector potential be written in the gauge A(r) = B z eˆx

(1854)

and the scalar potential is also a function of z φ(z) = − | E | z 33 L.D.

Landau, Z. f. Physik, 64, 629 (1930).

400

(1855)

The Hamiltonian is given by  2  e ˆ = 1 H pˆx − Bz + pˆ2y + pˆ2z − e | E | z 2m c

(1856)

which only depends on z. The momentum components pˆx and pˆy are good quantum numbers. The eigenfunctions can be written as   Φkx ,ky ,n (r) = exp i ( kx x + ky y ) φn (z) (1857) and φn (z) satisfies    2    h2 ky2 ¯ 1 e φn (z) = ¯h kx − B z + pˆ2y + pˆ2z − e | E | z φn (z) E− 2m 2m c (1858) This is recognized as the eigenvalue equation for a linear harmonic oscillator centered on z0 c |E| + m c2 (1859) z0 = h ¯ kx eB e B2 Thus, the energy eigenvalue is given by E = h ¯ ωc ( n +

¯h2 ky2 1 | E |2 |E| ) + − m c2 − ¯h kx c 2 2 2m 2B B

(1860)

The group velocity is given by vx

=

∂E ∂px

= −c

|E| B

(1861)

——————————————————————————————————

4.7.4

The Degeneracy of the Landau Levels

We shall consider a particle in a uniform magnetic field aligned along the z direction. The particle is confined to move within a large volume. Let the volume be bounded by surfaces, so that the accessible volume is described by Lx > x > 0 Ly > y > 0 Lz > z > 0

(1862)

and the wave function should be zero outside this volume. Instead of applying these boundary conditions, we shall impose Born-von Karman or periodic 401

boundary conditions. The vector potential A(r) is defined to be the solution of B = ∇ ∧ A(r)

(1863)

A(r) = − y Bz eˆx

(1864)

One solution is given by which corresponds to a particular choice of the gauge. The Hamiltonian of the charged particle in the field can be written as  2   q 1 1 ˆ = H pˆx − Ax (r) + pˆ2y + pˆ2z 2m c 2m  2   1 q 1 2 2 (1865) pˆx + y Bz + pˆy + pˆz = 2m c 2m in the region where the confining potential is zero. The wave function that satisfies the boundary conditions, can be written as     1 2 π nx x 2 π nz z Ψ(r) = √ exp i exp i f (y) (1866) Lx Lz Lx Lz where nx and nz are integers, nx = 0 , ± 1 , ± 2 , . . . etc. It is then found that the function f (y) satisfies the eigenvalue equation # "  2   2π¯ h nx q ¯h2 ∂ 2 2 π2 h ¯ 2 n2z 1 + y Bz − f (y) = E − f (y) 2m Lx c 2 m ∂y 2 m L2z (1867) or "  2 #   h2 ∂ 2 ¯ q 2 Bz2 2 π ¯h c nx 2 π 2 ¯h2 n2z − + + y f (y) = E − f (y) 2 m ∂y 2 2 m c2 q Bz Lx m L2z (1868) which can be re-written in the form of a displaced harmonic oscillator equation, by writing q Bz ωc = (1869) mc Then, the eigenvalue equation becomes "  2 #   h2 ∂ 2 ¯ m ωc2 nx Ly Φ0 2 π 2 ¯h2 n2z − + y + f (y) = E − f (y) 2 m ∂y 2 2 Φ m L2z (1870) where we have introduced the flux Φ, defined by Φ = Lx Ly Bz 402

(1871)

and the fundamental flux quantum Φ0 defined by Φ0 =

2 π ¯h c q

(1872)

Therefore, the eigenvalue E is just given by E = h ¯ ωc ( n +

1 ) 2

(1873)

and the eigenfunctions f (y) are the harmonic oscillator wave functions, φn (y) shifted through − nx ΦΦ0 Ly . That is, the normalized wave function is given by   Φ0 f (y) = φn y + nx Ly (1874) Φ

The degeneracy can be found, in the semi-classical approximation, by assuming that the deviation of the position of the harmonic oscillator from the equilibrium value y0 = − nx Ly ΦΦ0 is minimal. Under these circumstances, one has 0 < y0 < Ly . Hence, nx is restricted to be such that Ly > − nx Ly

Φ0 > 0 Φ

(1875)

Hence, the degeneracy is the given by the maximum number of values that nx can take. The degeneracy N is given by   Φ N = (1876) Φ0 which is controlled by the number of flux quanta, threading through the area Lx Ly . ——————————————————————————————————

4.7.5

Exercise 103

Find the energy eigenvalues and eigenfunctions of a particle of charge q and mass m moving in two dimensions, in the presence of a uniform magnetic field. Calculate the degeneracy of the lowest Landau level. Use the symmetric gauge where Φ A = r eˆϕ (1877) 2 π R02 and Φ is the total flux enclosed in the circular area π R02 . ——————————————————————————————————

403

4.7.6

Solution 103

The energy eigenvalue equation is     2  h2 1 ∂ ¯ ∂φ 1 ¯h ∂ q Φ − r + − i − r φ = Eφ 2 m r ∂r ∂r 2m r ∂ϕ c 2 π R02 (1878) ˆ z commutes with the Hamiltonian, The z component of the angular momentum L and so one can find simultaneous eigenfunctions which are of the form   1 exp i µ ϕ R(r) (1879) φ(r, ϕ) = √ 2π Hence, the radial wave function R(r) is given by     2  h2 1 ∂ ¯ ∂R 1 ¯h µ q Φ − r + − r R = E R (1880) 2 m r ∂r ∂r 2m r c 2 π R02 The radial wave function must satisfy boundary conditions at r r → ∞. The physically acceptable form of the solution is   µ R(r) = ρ 2 exp − ρ f (ρ) where ρ =

= 0 and

qΦ r2 c 4 π ¯h R02

(1881)

(1882)

and f (ρ) is a finite order polynomial. The derivatives w.r.t. r can be expressed as derivatives with respect to ρ via ∂ ∂r

= =

∂ρ ∂ ∂r ∂ρ ∂ qΦ r c 2 π ¯h R02 ∂ρ

(1883)

and ∂2 ∂r2

  2 2 qΦ ∂ qΦ 2 ∂ = + r c 2 π ¯h R02 ∂ρ c 2 π ¯h R02 ∂ρ2    2 qΦ ∂ ∂ = + 2ρ 2 c 2 π ¯h R0 ∂ρ ∂ρ2 

Thus, the eigenvalue equation becomes   2  ¯h q Φ ∂2R ∂R 1 µ − ρ − + − ρ R = ER m c 2 π R02 ∂ρ2 ∂ρ ρ 2

(1884)

(1885)

On substituting the form R(r) = ρ

µ 2

 exp

404

 − ρ

f (ρ)

(1886)

into the differential equation, we find      hqΦ ¯ ∂2 ∂f E m c 2 π R02 −ρ 2 +(2ρ−µ−1) + 1− f = 0 m c 2 π R02 ∂ρ ∂ρ ¯h q Φ (1887) This differential equation can be solved by the Fr¨obenius method, by expanding the solution in the form n=n Xr f (ρ) = an ρn (1888) n=0

On substituting the series into the differential equation, one finds the recursion relation   E m c 2 π R02 an (1889) an+1 ( n + 1 ) ( n + µ + 1 ) = 2n + 1 − ¯h q Φ The series for the polynomial f (ρ) truncates if the energy is given by   ¯ qΦ h ( 2 nr + 1 ) E = m c 2 π R02

(1890)

where nr is an integer. In this case the boundary conditions are satisfied, and f (ρ) is a polynomial of degree nr . The lowest energy eigenstates corresponds to nr = 0 and the eigenfunction can be written as     qΦ µ 2 r φ(r, ϕ) ∝ r exp i µ ϕ exp − c 4 π ¯h R02   qΦ µ 2 2 ∝ ( x + i y ) exp − (x + y ) c 4 π ¯h R2     02  Φ x + y2 µ ∝ ( x + i y ) exp − (1891) Φ0 2 R02 The probability density is peaked on a circle around the origin. The radius of the circle, rm , is given by 2 rm

c 2 π ¯h R02 qΦ   Φ0 = µ R02 Φ = µ

(1892)

All the states with different µ are degenerate. The degeneracy is given by the maximum number of different allowed µ values. For a disk of area π R02 the allowed values of rµ satisfy rµ < R0 , thus, the degeneracy is given by   Φ µmax = (1893) Φ0 ——————————————————————————————————

405

4.7.7

The Aharonov-Bohm Effect

The Aharonov-Bohm effect34 provides a conclusive demonstration that a charged quantum mechanical particle is sensitive to the vector potential, and not to the magnetic induction field B. In the Aharonov-Bohm effect, a charged particle moves in a region of space where the magnetic field is zero. The magnetic field free region is multiply connected, as a magnetic field threads through regions where the particle is excluded from35 . In order to provide a simple example of the Aharonov-Bohm effect, consider a charged particle which is confined to move on ring of radius R. A magnetic flux of strength Φ threads through the center of the ring and is contained within a cylinder of radius a.

Flux contained in a cylinder of radius a a

B

R

Figure 102: In the version of the Aharonov-Bohm effect being considered, the applied magnetic flux Φ is confined within a cylinder of radius a. The charged particle is confined to move on a ring of radius R which encircles the cylinder. The magnetic induction is in the z direction, so B = eˆz Bz . The magnetic flux through the center of the ring is given by Z Φ = dx dy Bz 34 Y.

Aharonov and D. Bohm, Phys. Rev. 115, 485 (1959). Tonomura, N. Osakabe, T. Matsuda, T. Kawasaki, J. Endo, S. Yano, and H. Yamada, Phys. Rev. Lett. 56, 792 (1986). 35 A.

406

Z =

d2 S . B

(1894)

in which the integral runs over the area inside the ring, where B is finite. Since the magnetic field is related to the vector potential via B = ∇ ∧ A

(1895)

then the flux is given by Z Φ =

2



d S.

 ∇ ∧ A

On using Stoke’s theorem, one finds the total flux is given by I Φ = dr . A

(1896)

(1897)

where the integral is around the perimeter of a ring of radius r and r > a. If the tangential component of the vector potential is denoted by Aϕ , then the loop integral is given by Φ = 2 π r Aϕ (1898) Hence, on considering the symmetry of the problem, one finds the vector potential is given by Φ A = eˆϕ (1899) 2πr There is no magnetic field present at this radius, since the definition B = ∇ ∧ A

(1900)

yields the z-component as Bz

∂ 1 ∂ Φ ) ∧ eˆϕ + eˆϕ ∂r r ∂ϕ 2πr   ∂ Φ 1 ∂ˆ eϕ Φ + eˆϕ ∧ = eˆr ∧ eˆϕ ∂r 2 π r r ∂ϕ 2 π r   ∂ Φ 1 Φ = eˆr ∧ eˆϕ − eˆϕ ∧ eˆr ∂r 2 π r r 2πr = 0 =

( eˆr

(1901)

The magnetic induction is zero since the total flux Φ contained within a loop is constant, if r ≥ a. The vector potential in the region where the magnetic field is finite 0 < r < a is given by A =

rΦ eˆϕ 2 π a2

407

(1902)

The energy eigenvalue equation for the particle confined on the ring of radius R is  2 1 ¯h ∂ qΦ − i − Ψ(ϕ) = E Ψ(ϕ) (1903) 2m R ∂ϕ c2πR This has normalized eigenfunctions given by 1 exp Ψ(ϕ) = √ 2πR



 iµϕ

where the energy eigenvalue E is determined by µ from  2 1 qΦ µ h ¯ − = E 2 m R2 2πc

(1904)

(1905)

Thus, the energy is given by ¯h2 2 m R2



qΦ µ − 2 π ¯h c

2 = E

(1906)

where 2 πqh¯ c = Φ0 is the fundamental flux quantum. The modulus of the wave function must be single valued, therefore Ψ(ϕ) = Ψ(ϕ + 2π) or

 1 = exp

(1907)

 iµ2π

(1908)

Hence, µ must be an integer, m = 0 , ± 1 , ± 2 . . . etc. The energy is given by  2 ¯h2 Φ E = m − (1909) 2 m R2 Φ0 and the ground state energy and wave function is determined by theinteger  Φ value of m which minimizes E and, therefore, depends on the ratio of . Φ0 The ground state corresponds to the value of m such that   m − 1 Φ m + 1 < < 2 Φ0 2

(1910)

Thus, the energy depends on the vector potential but not on the magnetic field B in the region where the particle moves. Furthermore, the ground state energy is a periodic function of Φ with periodicity Φ0 , as the value m changes discontinuously from m to m + 1 as Φ increases by an amount Φ0 . The current I produced by the charged particle flowing around the ring is defined classically as ∂E I = c (1911) ∂Φ 408

Energy eigenvalues 0.4

E(Φ Φ)

0.3

0.2

0.1

0 -4

-3

-2

-1

0

1

2

3

4

Φ/Φ0

Figure 103: The energy eigenvalues Em in the Aharonov-Bohm effect, as a function of the applied magnetic flux Φ threading the ring. The quantum mechanical current operator is defined as ˆ ∂H Iˆ = c ∂Φ   q ¯h ∂ Φ = i + 2 π m R2 ∂ϕ Φ0

(1912)

Thus, in the ground state the current has an expectation value I , given by   q ¯h Φ I = − m − (1913) 2 π m R2 Φ0 The current jumps discontinuously as the flux through the loop is increased by Φ0 . —————————————————————————————————— Example Find the energy eigenvalues and eigenfunctions of a particle of charge q and mass m moving in two dimensions, in the presence of a uniform magnetic field.

409

Flux induced current 0.75 0.5

I(Φ Φ)/I0

0.25 0 -4

-3

-2

-1

0

1

2

3

4

-0.25 -0.5 -0.75

Φ/Φ0

Figure 104: The dependence of the current I in the ground state on the applied flux Φ. Use the circularly symmetric gauge where Φ eˆϕ 2πr where Φ is the magnetic flux penetrating a ring of radius R0  r. A =

(1914)

—————————————————————————————————— Solution



The energy eigenvalue equation is    2  h2 1 ∂ ¯ ∂φ 1 ¯h ∂ q Φ −i − φ = E φ (1915) − r + 2 m r ∂r ∂r 2m r ∂ϕ c 2πr

ˆ z commutes with the Hamiltonian, The z component of the angular momentum L and so one can find simultaneous eigenfunctions which are of the form   1 φ(r, ϕ) = √ exp i µ ϕ R(r) (1916) 2π Hence, the radial wave function R(r) is given by     2  h2 1 ∂ ¯ ∂R 1 q Φ − r + ¯h µ − R = E R (1917) 2 m r ∂r ∂r 2 m r2 c 2π 410

The radial wave function must satisfy boundary conditions at r = 0 and r → ∞. The acceptable form for the solution at r = R0 → 0 vanishes at the origin as R(r) = rα (1918) which minimizes the effect of the centrifugal potential. On introducing a dimensionless variable ρ = kr (1919) where

¯ 2 k2 h 2m and the elemental flux quantum Φ0 , defined by

(1920)

E =

Φ0 =

c 2 π ¯h q

(1921)

one finds that the equation can be put in the dimensionless form     2  ∂R 1 Φ 1 ∂ ρ + 1 − 2 µ − R = 0 ρ ∂ρ ∂ρ ρ Φ0

(1922)

This is Bessel’s equation. The physically acceptable solution is finite in the limit as r → 0 and, therefore, the required solution is proportional to the Bessel function of order ν where   Φ ν = µ − (1923) Φ0 The eigenfunction is given by 1 φ(r, ϕ) = √ exp 2π



 iµϕ

Jν (kr)

(1924)

The Bessel function has the asymptotic variation Jν (ρ) ∼ ρν

(1925)

as ρ → 0 and varies as r Jν (ρ) ∼

2 cos πρ

 ρ − (ν +

1 π ) 2 2

 (1926)

when ρ → ∞. Note that, if the flux quantum threading the origin is changed so that ν increases, the maxima of the probability density moves radially outward, however, there is no change in energy. The absence of any energy change is not surprising in view of the fact that B = 0 in the region where the particle is moving and so, the effect of the vector potential can be absorbed into the phase of the wave function by a Gauge transformation. ——————————————————————————————————

411

4.8

The Pauli Spin Matrices

The Pauli Matrices are three two by two matrices, which have commutation relations similar to the commutation relations of the three components of the angular momentum operators. The Pauli matrices are represented by ! 0 1 σx = (1927) 1 0 −i 0

!

σy =

0 i

0 −1

!

σz =

1 0

(1928)

(1929)

These matrices are traceless, and have the properties that their square is equal to the unit matrix, σ0 , σx2 = σy2 = σz2 = σ0 (1930) where the unit matrix is given by 1 0

σ0 =

0 1

! (1931)

The three Pauli matrices and the unit matrix are linearly independent, and form a basis for the two by two matrices. That is, any two by two matrix can be expressed as a linear combination of the identity and Pauli matrices. Furthermore, they also satisfy the commutation relations [ σ x , σy ] = 2 i σ z [ σ z , σx ] = 2 i σ y [ σ y , σz ] = 2 i σx

(1932)

When the Pauli matrices are multiplied by a factor of h¯2 , these commutation relations become identical to the commutation relations of the components of angular momentum, for fixed l. Thus, if one multiplies the Pauli matrices by h ¯ 2 , one can identify these matrices as representing angular momentum operators where the total angular momentum corresponds to l = 1/2. The spin angular momentum operators are defined in terms of the Pauli matrices via ¯ h Sˆi = σi 2

(1933)

where the index i can take on the values x, y and z. The Pauli spin operators act on the space of two-component column vectors, Ψ, Ψ =

Ψ+ Ψ− 412

! (1934)

An arbitrary operator Aˆ given by Aˆ =

A1,1 A2,1

A1,2 A2,2

! (1935)

transforms the state Ψ according to the laws of matrix multiplication. That is, Aˆ acting on the state Ψ produces another state given by ! ! A1,1 A1,2 Ψ+ ˆ AΨ = A2,1 A2,2 Ψ− ! A1,1 Ψ+ + A1,2 Ψ− = (1936) A2,1 Ψ+ + A2,2 Ψ− A vector Ψ† which is the dual of the column vector Ψ is defined as the complex conjugate of the row vector ! Ψ† =

Ψ∗+

Ψ∗−

(1937)

An inner product can be defined for any two vectors on this space as the complex number formed from the components of the vectors ! ! Ψ + Φ∗+ Φ∗− Φ† Ψ = Ψ−   = Φ∗+ Ψ+ + Φ∗− Ψ− (1938) which, again, uses the laws of matrix multiplication. Physical states Ψ are to be normalized such that Ψ† Ψ = 1 (1939) which results in the normalization condition   | Ψ+ |2 + | Ψ− |2 = 1

(1940)

The existence of the inner product allows one to define the adjoint or Hermitean conjugate of the operator Aˆ as the operator Aˆ† which has the effect ! ! ! A1,1 A1,2 Ψ+ † ˆ ∗ ∗ Φ+ Φ− Φ AΨ = A2,1 A2,2 Ψ− ! ! ! A1,1 A2,1 Φ∗+ Ψ+ Ψ− = A1,2 A2,2 Φ∗−  ∗ † ˆ† = Ψ A Φ (1941)

413

Hence, the Hermitean conjugate of Aˆ is given by the transpose of the complex conjugate of Aˆ ! ∗ ∗ A A † 1,1 1,2 Aˆ = (1942) A∗2,1 A∗2,2 A Hermitean matrix is a matrix for which Aˆ = Aˆ† . The Pauli spin operators are Hermitean, and hence can represent physical quantities such as the components of the angular momentum. The vector spin angular momentum is defined as S

¯ h σ 2 ¯h ¯h ¯h = eˆx σx + eˆy σy + eˆz σz 2 2 2 =

(1943)

and the magnitude is given by c2 S

= Sˆx2 + Sˆy2 + Sˆz2 = =

¯2 2 h ¯h 2 2 ¯2 2 h σx + σy + σz 4 4 4 3 2 ¯h σ0 4

(1944)

The unit matrix σ0 is the identity operator as, when it acts on an arbitrary state, Ψ, it has the effect σ0 Ψ = Ψ (1945) and so the eigenvalues of the unit matrix are unity. As the eigenvalues of the magnitude of the spin angular momentum are s ( s + 1 ) ¯h2 , this means that s = 12 . The Pauli spin operators only act on the space formed by s = 12 . c2 have a degeneracy of This space is two-dimensional as the eigenvalues of S ( 2 s + 1 ) = 2. Thus, there are two independent basis states which can be chosen as the two column vectors that are the eigenvectors of σz σz χ± = ± χ±

(1946)

The eigenvector corresponding to the up-spin state with spin eigenvalue of + h¯2 is ! 1 χ+ = (1947) 0 while the eigenvector corresponding to the down-spin state, with spin eigenvalue − h¯2 is represented by ! 0 1

χ− =

414

(1948)

It should be noted that any pair of Pauli matrices, with different indices i and j, anti-commute. That is σi σj + σj σi = 2 σ0 δi,j

(1949)

The raising and lowering operators are defined as σ ± = σ x ± i σy

(1950)

which yields the representations 2 0

!

σ+ =

0 0

0 0

!

σ− =

0 2

and

(1951)

(1952)

——————————————————————————————————

4.8.1

Exercise 104

Prove that the above two column vectors of eqn(1947) and eqn(1948) are eigenstates of the z component of the spin operator Sz = h¯2 σz , and find the effect of the raising and lowering operators on these states. ——————————————————————————————————

4.8.2

Solution 104

The column vector χ+ is an eigenvalue of σz , since it satisfies the eigenvalue equation ! ! 1 0 1 σz χ+ = 0 −1 0 ! 1 = 0 = χ+

(1953)

with eigenvalue 1. Also, the column vector χ− is an eigenvalue of σz , since it satisfies the eigenvalue equation ! ! 1 0 0 σz χ− = 0 −1 1 415

=

0 −1

!

0 1

!

= − = − χ−

(1954)

with eigenvalue − 1. Hence, χ± are eigenstates of Sˆz with eigenvalues of ±

h ¯ 2.

The effect of the raising operator, Sˆ+ = Sˆx + i Sˆy on the eigenstates of Sˆz with eigenvalue − h¯2 is found from the effect of σ+   ¯h ¯h σ x + i σy (1955) Sˆ+ = σ+ = 2 2 acting on the eigenstates. This leads to σ+ χ−

=

0 0

=

2 0

=

2 χ+

2 0

!

0 1

!

!

(1956)

This shows that the raising operator increases the eigenvalue from − When the raising operator acts on the state χ+ it has the effect ! ! 0 2 1 σ+ χ+ = 0 0 0 ! 0 = 0 = 0

h ¯ 2

to +

h ¯ 2.

(1957)

The raising operator annihilates the state with the maximal eigenvalue of Sˆz , since by definition it cannot raise it any further. The effect of the lowering operator, σ− = σx − i σy on the eigenstates of ˆ Sz are given by ! ! 0 0 1 σ− χ+ = 2 0 0 ! 0 = 2 (1958) 1

416

which lowers the eigenvalue. However, since the state χ− has the lowest eigenvalue of σz one finds ! ! 0 0 0 σ− χ− = 2 0 1 ! 0 = 0 = 0

(1959)

Thus, the lowering operator produces zero when it acts on the state with the lowest eigenvalue. These equations can be summarized as Sˆ+ Sˆ+ Sˆ− Sˆ−

χ− χ+ χ+ χ−

= h ¯ χ+ = 0 = h ¯ χ− = 0

(1960)

——————————————————————————————————

4.8.3

Exercise 105

Show that any two by two matrix can be expressed as a linear combination of the Pauli matrices and the unit matrix. Also show that any two-component column vector Ψ can be represented as the linear superposition of the two eigenstates of σz , where ! Ψ+ Ψ = (1961) Ψ−

——————————————————————————————————

4.8.4

Solution 105

Consider an arbitrary two by two matrix A C

B D

417

! (1962)

This can be expressed as the sum of four two by two matrices each with just one non-zero element ! ! ! A B 1 0 0 1 = A + B C D 0 0 0 0 ! ! 0 0 0 0 +C + D 1 0 0 1 (1963) The four by four matrices with components on the diagonal can be expressed in terms of the sum of the unit matrix σ0 and σz via ! " ! !# 1 1 0 1 0 1 0 = + (1964) 0 1 0 −1 0 0 2 while the matrix 0 0

0 1

!

1 = 2

"

1 0

0 1

! −

1 0 0 −1

!# (1965)

can be expressed as the difference. The off-diagonal matrices can be expressed in terms of σx and σy via ! " ! !# 1 0 1 0 1 0 −i = + i (1966) 0 0 1 0 i 0 2 while the matrix 0 0 1 0

!

1 = 2

"

0 1

1 0

! − i

0 −i i 0

!# (1967)

can be expressed as the difference. Hence, we have the result that the arbitrary two by two matrix can be expressed as ! 1 1 1 i A B = ( A + D ) σ0 + ( A − D ) σz + ( B + C ) σx + ( B − C ) σy C D 2 2 2 2 (1968) This proves that the Pauli matrices and the identity span the linear space formed by two by two matrices. The states χ+ and χ− are eigenstates of σz . Since σz is a Hermitean operator, the eigenstates form a complete set. Thus, χ+ and χ− spans the space of the 418

column vectors as an arbitrary vector can be expressed as ! Ψ+ Ψ = Ψ− ! ! Ψ+ 0 = + 0 Ψ− ! ! 1 0 = Ψ+ + Ψ− 0 1 =

Ψ+ χ+ + Ψ− χ−

(1969)

as was expected. ——————————————————————————————————

4.8.5

Exercise 106

Prove the identity        σ.a σ.b = σ0 a . b + iσ. a ∧ b

(1970)

in which the spin vector is written in terms of the Pauli matrices as σ = eˆx σx + eˆy σy + eˆz σz

(1971)

——————————————————————————————————

4.8.6

Solution 106

First we note that the Pauli matrices satisfy the commutation relations X [ σ i , σj ] = 2 i i,j,k σk (1972) k

where i,j,k is the anti-symmetric Levi-Civita symbol. Also we note that the Pauli matrices satisfy σ i σ j = − σ j σi

(1973)

if i 6= j, whereas if i = j then σ i σi = σ 0 419

(1974)

where σ0 is the unit matrix. These relations can be combined in the anticommutation relation { σi , σj }+ = 2 δi,j σ0 (1975) ˆ is defined as { Aˆ , B ˆ }+ , in which the anti-commutator of two operators Aˆ and B where ˆ }+ = Aˆ B ˆ + B ˆ Aˆ { Aˆ , B (1976)

Hence, on adding the commutation and anti-commutation relations, one has the identity X σi σj = δi,j σ0 + i i,j,k σk (1977) k

The identity to be proved follows immediately by expressing the scalar product in terms of the sum of Cartesian components    X X σ.a σ.b = σi ai σ j bj i

=

j

X

σi σj ai bj

i,j

=

X 

δi,j σ0 + i

i,j

=



X

i,j,k σk

X

δi,j σ0 ai bj + i

X

i,j

σk



 a.b

+ i

X

X

 σk

 a ∧ b k

k



 a.b

 + iσ.

i,j,k ai bj

i,j

k

= σ0 = σ0

ai bj

k

 a ∧ b

(1978)

This completes the proof. ——————————————————————————————————

4.8.7

Exercise 107

Find all the eigenfunctions and eigenvalues for the Pauli matrix σx , and show that they form a complete orthonormal set. ——————————————————————————————————

420

4.8.8

Solution 107

The matrix σx is given by σx =

0 1

1 0

! (1979)

and the eigenfunctions satisfy the eigenvalue equation σx χx = λ χx

(1980)

The eigenfunctions are represented by column vectors of the form ! α x χ = β

(1981)

which are normalized when | α |2 + | β |2 = 1

(1982)

The eigenvalue equations are expressed as the coupled algebraic equations for the components of Ψ as ! ! ! 0 1 α α = λ (1983) 1 0 β β or −λ 1

1 −λ

!

α β

! = 0

(1984)

This equation only has a solution if the inverse matrix does not exist. Since the inverse matrix is inversely proportional to the determinant, the determinant must vanish for there to be a non-trivial solution for the components. Therefore, the eigenvalue λ is given by the solution of 1 −λ (1985) = 0 −λ 1 or λ2 = 1

(1986)

which is consistent with the relation σx2 = σ0

(1987)

The eigenvalues of σx , are given by λ = ±1

421

(1988)

The eigenfunction χx+1 corresponding to λ = 1 is found from ! ! ! 0 1 α α = 1 0 β β which yields α = β. Thus, the normalized eigenfunction is given by ! 1 1 x χ+1 = √ 1 2

(1989)

(1990)

up to an arbitrary phase. The eigenfunction corresponding to λ = − 1 is found from ! ! ! 0 1 α α = − 1 0 β β

(1991)

which yields α = − β. Thus, the normalized eigenfunction is given by ! 1 1 χx−1 = √ (1992) −1 2 up to an arbitrary phase. The eigenstates of σx are orthogonal as χx+1 † χx−1

=

1 2 = 0

=

!

1 2

1

1



1 −1

!

 1 − 1 (1993)

Furthermore, since the space of column vectors is two-dimensional, the set of eigenfunctions are complete. That is, any arbitrary vector   α Ψ = (1994) β can be expressed as       (α − β) 1 (α + β) 1 1 1 α √ √ √ √ (1995) + = −1 1 β 2 2 2 2

422

where the expansion coefficients were determined from the inner products with the eigenstates (α + β) √ (1996) χx+1 † Ψ = 2 and χx−1 † Ψ =

(α − β) √ 2

(1997)

Thus, we have shown that an arbitrary state can be expanded in terms of the eigenstates of σx as Ψ =

(α + β) x (α − β) x √ √ χ+1 + χ−1 2 2

(1998)

—————————————————————————————————— Thus, the Pauli matrices form a representation of half-integer angular momentum, called spin. The spin quantum numbers are intrinsic to the particle and, therefore, the operators cannot be represented in terms of the position and momentum of the particle. The spin of a particle first became manifest in the anomalous Zeeman effect, in which there is a coupling between the spin state and an external magnetic field. ——————————————————————————————————

4.8.9

Exercise 108

Find the eigenvalues and eigenvectors of the component of the spin along the unit vector ηˆ, where ηˆ = sin θ cos ϕ eˆx + sin θ sin ϕ eˆy + cos θ eˆz

(1999)

Work in the basis in which σ ˆz is diagonal. Hence, find the unitary matrix Ueˆ(θ) which produces a rotation of the spin states through an angle θ, about the axis eˆ = − sin ϕ eˆx + cos ϕ eˆy

(2000)

perpendicular to the plane of eˆz and ηˆ. Hint: In addition to producing the transformation of the spin state “aligned” parallel to the axis z to the spin state “aligned” with ηˆ, it is also required that the rotation does not change the “direction” of spin states aligned parallel to the rotation axis. This is indicated schematically in figs(105) and (106) for a 423

Rotation of a set of Spin States z Rotation angle θ

η

Sη = 1 ε axis of rotn.

φ x

y Sη = 0

Sη = - 1

Figure 105: A spin state with eigenvalue of Sˆz equal to +¯ h is to be transformed to a spin state which an eigenstate of Sˆη , with eigenvalue of +¯h. spin one particle. ——————————————————————————————————

4.8.10

Solution 108

We seek the eigenvalues and eigenfunctions of the operator η . σ = sin θ cos ϕ σ ˆx + sin θ sin ϕ σ ˆy + cos θ σ ˆz which has the matrix representation   cos θ sin θ exp[ − i ϕ ] η.σ = h ¯ sin θ exp[ + i ϕ ] − cos θ

(2001)

(2002)

in the basis where σ ˆz is diagonal. The eigenvalues µ ¯h are determined from the secular equation cos θ − µ sin θ exp[ − i ϕ ] 0 = (2003) − cos θ − µ sin θ exp[ + i ϕ ]

424

Rotation of a set of Spin States z Sε = 0

η θ

φ

Sε = - 1 x

Sε = 1

ε y

Sε eigenstates are invariant

Figure 106: It is also required that a spin rotation around the axis eˆ does not change the eigenstates of Sˆe . and, thus, are found to be µ = ± ¯h. The eigenfunction χηµ corresponding to the eigenvalue µ, is expressed in terms of its components through  η  αµ η χµ = (2004) βµη The components for µ = + 1 are found from the eigenvalue equation      η  0 ( cos θ − 1 ) sin θ exp[ − i ϕ ] α+ = (2005) η 0 sin θ exp[ + i ϕ ] − ( cos θ + 1 ) β+ Thus, the normalized eigenfunction of the operator in eqn(2001) corresponding to the µ = 1 eigenvalue is found as   cos θ2 η χ+ = (2006) sin θ2 exp[ + i ϕ ] Likewise, one finds the eigenfunction corresponding to µ = − 1 is given by   − sin θ2 exp[ − i ϕ ] (2007) χη− = cos θ2

425

The rotation is about an axis in the plane perpendicular to the unit vectors eˆz and eˆη . The matrix Ueˆ(θ) is determined from its action on the two eigenstates of σ ˆz χηµ = Ueˆ(θ) χzµ (2008) Thus, we have   z   η   αµ αµ cos θ2 − sin θ2 exp[ − i ϕ ] = βµη βµz sin θ2 exp[ + i ϕ ] cos θ2 (2009) On recognizing that the component of the spin operator along the rotation axis is given by     0 − i exp[ − i ϕ ] eˆ . σ ˆ = (2010) + i exp[ + i ϕ ] 0 and that the square is the unit matrix 2

 eˆ . σ ˆ

 =

1 0

0 1

 = σ0

we find that the spin rotation operator may be written in the form   θ θ Ueˆ(θ) = cos σ ˆ0 − i sin eˆ . σ ˆ 2 2 Hence, the spin rotation operator is given by   θ Ueˆ(θ) = exp − i eˆ . σ ˆ 2   iθ = exp − eˆ . Sˆ ¯h

(2011)

(2012)

(2013)

where the spin is defined as ¯ h σ ˆ Sˆ = 2

(2014)

——————————————————————————————————

4.8.11

The Pauli Equation

The existence of spin was first inferred by Uhlenbeck and Goudschmit36 as a result of experiments where atoms were placed in a strong magnetic field and 36 G.E. Uhlenbeck and S.A. Goudsmit, Naturwissenschaften, 13, 953 (1925) and Nature, 117, 264 (1926).

426

the absorption lines were measured. In the ordinary Zeeman effect, the orbital angular momentum of a charged electron produces a magnetic moment which couples to the field and, therefore, raises the degeneracy of the electronic levels. This results in a splitting of the absorption lines, when the field is applied. The number of lines that are seen is equal to the degeneracy of the atomic levels in the absence of the field which are labeled by n and l. In general, the degeneracy is expected to be given by 2 l + 1 corresponding to the different m values. However, due to the presence of spin, there are more lines seen than predicted by the spin-less version of the Schr¨odinger equation. Even the l = 0 state is split into two levels corresponding to a degeneracy of 2. On introducing a spin operator S representing spin angular momentum and equating the degeneracy 2 S + 1 = 2, one finds that the spin must be half-integer S = 12 . The Pauli equation37 assumes that the wave function for a spin half particle has two components ! Ψ+ (r) Ψ(r) = (2015) Ψ− (r) and expresses the kinetic energy operator of the free particle in terms of the product of a Laplacian and the unit two by two matrix, p ˆ2 2m

(2016)

( pˆ . σ ) ( pˆ . σ ) 2m

(2017)

Tˆ = which can be re-written as Tˆ =

by using the identity        σ.a σ.b = σ0 a . b + iσ. a ∧ b

(2018)

with a = b = p ˆ. Thus, the time-dependent Schr¨odinger equation for the two-component (column vector) wave function Ψ(r) representing a free particle can be written in the form of the field-free Pauli equation   ∂ ¯h2 i¯ h Ψ(r, t) = − ( σ . ∇ )2 Ψ(r, t) (2019) ∂t 2m The corresponding equation for a charged particle in an electromagnetic field is obtained by using the minimum coupling assumption pµ → pµ − qc Aµ . The resulting equation is found as 2    q 1 ∂ + σ0 q φ(r, t) Ψ(r, t) A(r, t) ) σ . ( − i ¯h ∇ − i ¯h Ψ(r, t) = c 2m ∂t (2020) 37 W.

Pauli, Z. f. Physik, 43, 601 (1927).

427

which, on using the identity, becomes   2 ∂ 1 q i ¯h Ψ(r, t) = σ0 − i ¯h ∇ − A(r, t) ∂t 2m c   ¯h q − σ . ∇ ∧ A(r, t) + A(r, t) ∧ ∇ 2mc  + σ0 q φ(r, t) Ψ(r, t) (2021) Furthermore, since  ∇ ∧ A Ψ(r) = Ψ(r)

 ∇ ∧ A

− A ∧ ∇ Ψ(r)

(2022)

and B = ∇ ∧ A, one finds   2 ∂ 1 q i¯ h Ψ(r, t) = σ0 − i ¯h ∇ − A(r, t) ∂t 2m c  ¯h q σ . B(r, t) + σ0 q φ(r, t) Ψ(r, t) (2023) − 2mc Since the spin angular momentum S is given by h¯2 σ, one finds that the spin couples to the magnetic field giving rise to the anomalous Zeeman effect, but the relation between spin angular momentum and magnetic moment M is   q M = 2 S (2024) 2mc which is different from the relation between the orbital angular momentum L and the magnetic moment found in the ordinary Zeeman effect   q M = L (2025) 2mc The extra factor of 2 found for the spin half particle is known as the gyromagnetic ratio, g, and the magnitude of the factor 2 qmh¯ c is known as the Bohr magneton µB .

4.8.12

Spin Dynamics

We shall consider the dynamics of a particle with spin S = h¯2 that is localized at some point in space, when a time-independent magnetic field B is applied to the system. The direction of the field may be taken to define the direction of the z axis of our Cartesian coordinate system. We shall only be concerned with the dynamics of the spin and, therefore, suppress any mention of the coordinate dependence of the wave function.

428

The wave function is then represented as a linear superposition of spin-up and spin-down states ! Ψ+ (t) Ψ(t) = (2026) Ψ− (t) where Ψ± (t) are the time-dependent expansion coefficients. The Pauli equation is then given by the matrix equation i¯ h

∂ Ψ(t) = − µB Bz σz Ψ(t) ∂t

(2027)

where the Hamiltonian only consists of the Zeeman term. Since the energy eigenstates correspond to eigenstates of the z component of the spin, Sz = h¯2 σz , one may decompose the time-dependent wave function into energy eigenstates Ψ(t) = Ψ+ (t) χ+ + Ψ− (t) χ−

(2028)

On substituting this into the time-dependent Schr¨odinger equation, one finds that ∂ i¯ h Ψ± (t) = ∓ µB B Ψ± (t) (2029) ∂t which have the solutions   µB B t Ψ± (t) = exp ± i Ψ± (0) (2030) ¯h Hence, we have found the time dependence of our arbitrary initial spin wave function Ψ(0) as     µB B t µB B t Ψ(t) = exp + i Ψ+ (0) χ+ + exp − i Ψ− (0) χ− (2031) h ¯ ¯h If the initial state was known at t = 0 to be an eigenstate of, say Sx with the eigenvalue + h¯2 , then one has   1 χ+ + χ− (2032) Ψ(0) = √ 2 or 1 Ψ(0) = √ 2

1 1

! (2033)

Thus, we have found that the time dependence of the wave function at later times, t, is given by       µB B t 1 µB B t exp + i Ψ(t) = √ χ+ + exp − i χ− ¯h ¯h 2 (2034)

429

From this one can see that the probability of finding the particle in an eigenstate of Sx with eigenvalue + h¯2 is going to oscillate with time, as is the probability that the particle is in the Sx eigenstate with eigenvalue − h¯2 . This oscillation corresponds to the classical precession of a spin around the z axis, with a frequency given by ω = 2 µh¯B B , known as the Larmor precession frequency.

Μ

B

Figure 107: A classical magnetic moment M in an applied magnetic field B experiences a torque τ = M ∧ B and, hence, precesses around the direction of the field. ——————————————————————————————————

4.8.13

Exercise 109

Evaluate the time dependence of the probability that a spin 12 in a magnetic field B aligned along the z - axis is found in the eigenstates of Sy , if the initial state is an eigenstate of Sx with eigenvalue + h¯2 . ——————————————————————————————————

4.8.14

Solution 109

The initial state is an eigenstate of Sˆx with eigenvalue + can be represented as ! 1 1 Ψ(0) = √ 1 2

430

h ¯ 2.

The initial state (2035)

which evolves with time as 



it ˆ − H ¯h

Ψ(t) = exp

Ψ(0)

(2036)

where the Hamiltonian is given by ˆ = − µB B σz H

(2037)

The time-dependent state is given by 1 Ψ(t) = √ exp 2



i t µB B h ¯



!

1 0

1 + √ exp 2



i t µB B − ¯h



0 1

!

(2038) The matrix σy has eigenvalues of ± 1 and the eigenvalues are given by ! 1 1 y χ±1 = √ (2039) ±i 2 respectively. The state Ψ(t) is expanded in terms of the eigenstates χy± as Ψ(t) =

X

C± (t) χy±

(2040)

±

where C± (t) =

1 2



 exp

i t µB B ¯h



 ∓ i exp



i t µB B ¯h

Thus, we have the time-dependent probability given by   2 µB B t 1 y 1 ∓ sin P± (t) = 2 ¯h

 (2041)

(2042)

——————————————————————————————————

4.8.15

Exercise 110

A spin half nucleus is placed in a static magnetic field B0 aligned along the z axis and a smaller rotating magnetic field B1 in the x - y plane. The frequency of the a.c. field is ω. If the nucleus is initially pointing in the + z direction at t = 0, what is the probability that it will be aligned with the + z axis at later times?

431

Experiments were suggested and preformed by I.I. Rabi and co-workers38 which were based on calculations of the above type. ——————————————————————————————————

4.8.16

Solution 110

The spin vector can be written as !

α(t) β(t)

Ψ(t) =

(2043)

and the Schr¨ odinger equation is   ∂ i¯ h Ψ(t) = − µN B0 σz + B1 σx cos ωt − B1 σy sin ωt Ψ(t) ∂t (2044) or ∂ i Ψ(t) = − ∂t

µN B1 h ¯

µN B0 h ¯

!

µN B1 h ¯

exp[ + i ω t ] − µNh¯ B0

exp[ − i ω t ]

Ψ(t)

(2045) We should note that the sense of the rotation of B1 about the z-axis is such that the axis of rotation is anti-parallel to B0 . We are also implicitly assuming that µN is positive. The equations for the components of the spinors are       ∂ µN B0 µN B1 α(t) = − α(t) − exp + i ω t β(t) i ∂t h ¯ ¯h (2046) and i

∂ β(t) = + ∂t



µN B0 h ¯



 β(t) −

µN B1 ¯h



 exp

 − iωt

α(t) (2047)

On solving the first equation for β and substituting this into the second, one obtains a second order differential equation for α ∂2 ∂ α(t) − i ω α(t) + 2 ∂t ∂t



µN ¯h

2

( B02 + B12 ) − ω

µN B0 ¯h

 α(t) = 0 (2048)

38 I.I. Rabi, Phys. Rev. 51, 652 (1937), I.I. Rabi, J.R. Zacharias, S. Millman and P. Kusch, Phys. Rev. 53, 318 (1938).

432

This is solved by assuming that α(t) has the form   α(t) = α(0) exp i λ t

(2049)

The characteristic equation is a quadratic equation in λ which can be solved, yielding s 2  2 µN B0 ω g µN B1 ω λ± = + ± − + (2050) 2 ¯h 2 ¯h The general solution can be written as     α(t) = a+ exp i λ+ t + a− exp i λ− t

(2051)

The components of β can be written as        β(t) = b+ exp i λ+ t + b− exp i λ− t exp − i ω t (2052) However, b+ and b− are given in terms of the a’s via the original differential equations. Thus, b+ is linear in a+ and b− is linear in a− s 2  2   h ¯ ω h ¯ ω ( 2 − µN B0 ) ∓ − µ B + µ B N 0 N 1 2 b± = a± µN B1 (2053) To satisfy the initial conditions one must have α(0) = a+ + a− = 1 and β(0) = b+ + b− = 0. These conditions allows one to solve for α(t). The probability of finding the spin-up eigenvalue is given by | α(t) |2 and  " 2

| α(t) |

= 1 −

 µN B0

2

× sin

s

2

# µN B1 × 2  2 h ¯ω + µN B1 − 2

µN B0 ω − ¯h 2

2

 +

µN B1 ¯h

2 t

(2054)

As the spin one half wave function is normalized and has an overall phase factor, the spin state only depends on two variables. Therefore, the spin one half wave function can always be interpreted in terms of a classical state in which the spin has a definite direction. When the frequency ω is away from the resonance frequency   µN B0 ω = 2 (2055) ¯h 433

the spin precesses around the z-axis with frequency ω and the z component also makes makes small amplitude oscillations. That is, the spin motion is confined within a small ring parallel the equator of the unit sphere. However, for frequencies near the resonance frequency, the z component of the spin performs large oscillations and almost flips to the other pole with frequency  amplitude  2

µN B1 h ¯

.

——————————————————————————————————

4.8.17

The Berry Phase

In the presence of a magnetic field, an isolated spin will find an equilibrium state which is the ground state. In the ground state the spin is aligned parallel to the field. If the direction of the field is subsequently changed, sufficiently slowly, one expects that the direction of the spin will continue to align with the magnetic field at every instant of time. That is, one does not expect the slowly time varying field to cause the spin to make a transition to states of higher energy. This expectation is borne out by detailed calculations on systems of various spin magnitudes. However, although the spin state does follow the field direction, the spin wave function does acquire an additional phase. If the field is returned to the original direction, the additional phase factor is determined by the topology of the spin’s motion39 . The motion of the spin can be visualized by projecting the spin direction onto the unit sphere. The direction of the spin can be specified by specifying the polar coordinates (θ(t), ϕ(t)). As the field is changed, the spins direction will map out a path on the unit sphere. The additional phase, or Berry phase δ, acquired by the spin is determined by the solid angle Ω enclosed by the spin’s orbit. The Berry phase is determined by the solid angle and the magnitude of the spin, in units of h ¯ . The path of the spin is parameterized as θ(ϕ). Since the infinitesimal solid angle dΩ is given by dΩ = dϕ dθ sin θ

(2056)

on integrating from the pole θ = 0 to θ(ϕ), the solid angle enclosed by the infinitesimal wedge is dΩ = dϕ ( 1 − cos θ(ϕ) ) Then, the solid angle enclosed by the complete orbit is given by I Ω = dϕ ( 1 − cos θ(ϕ) ) where the integral over ϕ runs over 2 π. 39 M.V.

Berry, Proc. R. Soc. Lond. A 392, 45 (1984).

434

(2057)

(2058)



(θ(t),ϕ(t))

Figure 108: The Berry Phase is determined by the solid angle Ω swept out by the spins orbit on the unit sphere. The Berry phase can be illustrated by considering a spin one half in a magnetic field of constant magnitude, B, oriented along the direction (θ, ϕ). In this case, the Zeeman Hamiltonian is given by ˆZ H

= − µB ( B . σ )   = − µB B cos θ σz + sin θ ( sin ϕ σy + cos ϕ σx ) (2059)

which, on using the Pauli-spin matrices, can be expressed as   cos θ sin θ exp[− i ϕ ] ˆ Z = − µB B H sin θ exp[ + i ϕ ] − cos θ

(2060)

ˆ Z has an eigenstate with For fixed (θ, ϕ), the time-independent Hamiltonian H eigenfunction given by   cos θ2 (2061) χ+ = sin θ2 exp[ + i ϕ ] corresponding to the eigenvalue E0 = − µB B 435

(2062)

Thus, in this eigenstate the spin is aligned parallel to the applied field. For a static field one has the time-dependent wave function given by     µB B cos θ2 exp + i t (2063) χ+ (t) = sin θ2 exp[ + i ϕ ] ¯h where the time dependence is given purely by the exponential phase factor. If the direction of the field (θ(t), ϕ(t)) is changed very slowly, one expects the spin will adiabatically follow the field direction. That is, if the field is rotated sufficiently slowly, one does not expect the spin to make a transition to the state with energy E = + µB B where the spin is aligned anti-parallel to the field. However, the wave function may acquire a phase which is different from the time and energy dependent phase factor expected for a static field. This extra phase is the Berry phase δ, which can be calculated from the Schr¨odinger equation      ∂ α(t) cos θ(t) sin θ(t) exp[− i ϕ(t) ] α(t) i¯ h = − µB B β(t) sin θ(t) exp[ + i ϕ(t) ] − cos θ(t) β(t) ∂t (2064) We shall assume that the wave function takes the adiabatic form !      µB B α(t) cos θ(t) 2 exp + i t − δ(t) = β(t) ¯h sin θ(t) exp[ + i ϕ(t) ] 2 (2065) which instantaneously follows the direction of the field, but is also modified by the inclusion of the Berry phase δ(t). On substituting this ansatz into the Schr¨ odinger equation, one finds that the non-adiabatic terms satisfy !   0 ∂ϕ ∂δ cos θ(t) 2 − + exp[ + i ϕ(t) ] sin θ(t) ∂t ∂t sin θ(t) exp[ + i ϕ(t) ] 2 2 ! i ∂θ − sin θ(t) 2 = 2 ∂t cos θ(t) exp[ + i ϕ(t) ] 2 (2066) The above equation is then projected onto the adiabatic state by multiplying it by the row matrix   θ(t) (2067) cos θ(t) sin exp[ − i ϕ(t) ] 2 2 One finds that the derivative of θ w.r.t. t cancels and that the equation simplifies to ∂δ ∂ϕ θ − + sin2 = 0 (2068) ∂t ∂t 2

436

Hence, the Berry phase is given by integrating w.r.t. to t, Z t θ(t0 ) ∂ϕ δ(t) = dt0 0 sin2 ∂t 2 Z0 θ(ϕ) = dϕ sin2 2 Z 1 = dϕ ( 1 − cos θ(ϕ) ) 2

(2069)

On completing one orbit in spin space, the extra phase is given by δ =

1 Ω 2

as was claimed.

437

(2070)

4.9

Transformations and Invariance

The assumption underlying the application of quantum mechanics to physical systems is that the basic structure and predictions of experimental results on quantum mechanical systems are independent of the inertial coordinate system used. If we consider space-time which is governed by Galilean invariance, time being absolute and space being isotropic and governed by the laws of Euclidean geometry, we have non-relativistic quantum mechanics. Transformations between different coordinate systems (passive transformations) leave the physical system unaltered and merely provide relations between reference systems of different observers. Alternately, two different physical systems may be related to each other by a transformation of the physical position of the physical system and measuring devices. Under an active transformation such as a physical rotation of one system, the two systems may be brought into coincidence. Equivalent quantum mechanical systems can be transformed into each other, ˆ . The unitary via a unitary transformation associated with a unitary operator U operator is defined as an operator that has a Hermitean conjugate which is identical to the inverse operator ˆ† = U ˆ −1 U

(2071)

This has a consequence that the eigenvalues of a unitary operator are complex numbers of magnitude unity, and the eigenfunctions form an orthonormal set. ˆ −1 U ˆ between two This can be seen by considering the matrix elements of U different eigenfunctions φλ1 and φλ2 Z Z ˆ −1 U ˆ φλ (r) = λ∗ λ1 d3 r φ∗λ2 (r) φλ1 (r) d3 r φ∗λ2 (r) U 2 1 Z = d3 r φ∗λ2 (r) φλ1 (r) (2072) ˆ is unitary. Thus, we have since U   Z ∗ λ2 λ1 − 1 d3 r φ∗λ2 (r) φλ1 (r) = 0

(2073)

Hence, the eigenvalues have modulus unity as λ∗1 λ1 = 1

(2074)

and eigenfunctions corresponding to different eigenvalues are orthogonal. Physically equivalent states are related via unitary transformations. Under a unitary transformation the states are transformed according to ˆ Ψ(r) Ψ(r) → Ψ0 (r) = U

438

(2075)

which preserves the normalization of Ψ0 as Z Z ˆ† U ˆ Ψ(r) d3 r Ψ0∗ (r) Ψ0 (r) = d3 r Ψ∗ (r) U Z = d3 r Ψ∗ (r) Ψ(r) =

1

(2076)

The operators are transformed according to ˆ Aˆ U ˆ† Aˆ → Aˆ0 = U

(2077)

as the expectation values of the transformed operators sandwiched between the transformed states leads to the same expectation values as the untransformed operators in the untransformed states Z Z d3 r Ψ0∗ (r) Aˆ0 Ψ0 (r) = d3 r Ψ∗ (r) Aˆ Ψ(r) (2078)

4.9.1

Time Translational Invariance

ˆ (t, t0 ) is a unitary operator which is defined The time translational operator U by its action on a wave function ˆ (t, t0 ) Ψ(r, t0 ) = Ψ(r, t) U

(2079)

Since the wave function Ψ(r, t) has the same normalization as Ψ(r, t0 ), the time evolution operator is a unitary operator as it is norm conserving. The time translational operator can be expressed in terms of the time derivative, since   ∂ exp ( t − t0 ) Ψ(r, t0 ) = Ψ(r, t) (2080) ∂t0 Thus, we find that the operator ∂ (2081) ∂t0 is the infinitesimal generator of time translations and the time translational operator can be expressed as   ∂ ˆ U (t, t0 ) = exp ( t − t0 ) (2082) ∂t0 The time dependence of the Schr¨odinger equation is governed by the Hamiltoˆ and if this is time independent then nian H, i ¯h

∂ ˆ ≡ H ∂t 439

(2083)

which gives the explicit term for the time translational operator   ˆ (t, t0 ) = exp − i ( t − t0 ) H ˆ U ¯h

(2084)

The time translational operator only depends on the time difference t − t0 . The time translation operators can be compounded, thereby forming a group operation ˆ (t2 − t1 ) U ˆ (t1 − t0 ) = U ˆ (t2 − t0 ) U (2085) ˆ only depends on the time differences t2 − t1 . The identity operator is since U ˆ( 0 ) = 1 U

(2086)

ˆ −1 ( t ) = U ˆ( − t ) U

(2087)

ˆ( t ) U ˆ( − t ) = U ˆ( 0 ) U

(2088)

ˆ ( t ) is The inverse of U such that

If a physical quantity is represented by a time-independent operator Aˆ in the Schr¨ odinger picture, and its expectation values are given by Z Z 3 ∗ ˆ ˆ † (t, t0 ) Aˆ U ˆ (t, t0 ) Ψ(r, t0 ) d r Ψ (r, t) A Ψ(r, t) = d3 r Ψ∗ (r, t0 ) U (2089) this is independent of time if ˆ † (t, t0 ) Aˆ U ˆ (t, t0 ) = Aˆ U

(2090)

ˆ which is satisfied if the operator Aˆ commutes with the Hamiltonian H, ˆ , Aˆ ] = 0 [H

(2091)

In particular, since time is homogeneous, and hence the Hamiltonian of a ˆ commutes closed system is independent of time, the energy is conserved as H with itself.

4.9.2

Translational Invariance

A translation of a system by a distance a changes the wave function of the system ˆ Ψ(r) → Ψ0 (r) = S(a) Ψ(r) (2092)

440

The unitary operator that represents a displacement of a system from r to r + a is given by   ˆ S(a) = exp − a . ∇ (2093) so ˆ = S(a) Ψ(r)   = exp − a . ∇ Ψ(r)

Ψ0 (r)

=

Ψ(r − a)

(2094)

Thus, the wave function Ψ0 evaluated at r +a has the same value as Ψ evaluated ˆ can be combined to yield other operators at r. The translational operators S(a) of the same type ˆ ˆ ˆ + b) S(a) S(b) = S(a (2095) The translation operators form a representation of a group. Abelian and it is a continuous group.

The group is

Translation through distance a 1

Ψ(x)

Ψ(x-a)

Ψ'(x)=Ψ(T-1x)=Ψ(x-a)

0.8

0.6

0.4

0.2

0

x'=Tx=x+a x0 0

1

x0+a 2

3

0

1

2

3

x

x

Figure 109: The translation of an arbitrary state Ψ(r) through a distance a. For an infinitesimally small displacement ξ, one has   ˆ S(ξ) = 1 − i ξ . − i∇ + O(ξ 2 ) =

1 −

i ξ . pˆ + O(ξ 2 ) ¯h 441

(2096)

so − i ∇ is the generator of infinitesimal transformations. The generator of infinitesimal transformations corresponds to the momentum operator. If a system is invariant under displacements then the expectation values of physical operators in these states should be the same for Ψ as Ψ0 . In particular, the Hamiltonian is invariant under the transformation ˆ S(ξ) ˆ ˆ ˆ †H = H S(ξ)

(2097)

ˆ must commute From the infinitesimal translation, we find that p ˆ and H ˆ ] = 0 [p ˆ, H

(2098)

Hence, we find pˆ is a constant of motion. Thus, for translationally invariant systems the momentum is conserved. If we translate the state of the system and the measuring devices then, as space is homogeneous, this is equivalent to a passive transformation in which the system remains unchanged but is represented in a different coordinate system. The measuring devices in this new coordinate system are represented by the operators Cˆ 0 which are given in terms of the operators of the old coordinate system via ˆ Cˆ 0 = S(a) Cˆ Sˆ† (a) (2099) The momentum and coordinate operators are transformed accordingly as pˆ0

ˆ = S(a) p ˆ Sˆ† (a)     = exp − a . ∇ p ˆ exp + a . ∇ = pˆ

and r0

ˆ = S(a) r Sˆ† (a)     = exp − a . ∇ r exp + a . ∇ = r − a

(2100)

For an active transformation, in which only the system is translated and the measuring systems are kept in place, the corresponding operators are those of the original system. When the homogeneity of space is disturbed, for example by the application of a magnetic field, the translation operators take a different form. For a uniform magnetic field and no electrostatic potential, the system should be uniform and

442

so the translation operator should commute with the Hamiltonian. In this case, the correct generator of the translation is given by the pseudo-momentum q q ˆ = p K ˆ − A + B ∧ r c c

(2101)

where the sum of the first two terms is recognized as the canonical momentum. The translation through a distance a is given by   i ˆ ˆ S(a) = exp − a.K (2102) ¯h The components of the pseudo-momenta do not commute, as ˆi , K ˆ j ] = − i ¯h q i,j,k Bk [K c and so the translation operators also do not commute   q ˆ ˆ ˆ ˆ S(a) S(b) = S(b) S(a) exp + i B.(a ∧ b) ¯h c

(2103)

(2104)

The non-commutivity of the translational operators is a manifestation of the presence of an Aharonov - Bohm phase.

4.9.3

Periodic Translational Invariance

Consider a potential V (x) which is periodic in translations through any integer multiple of a discrete repeat distance a. That is, V (x + na) = V (x)

(2105)

This potential is periodic in x with period a. We seek eigenvalues of the energy eigenvalue equation



2

pˆ 2m

ˆ φ(x) = E φ(x) H  + V (x) φ(x) = E φ(x)

ˆ The unitary operator S(na) defined by   ina ˆ S(na) = exp − pˆ ¯h

(2106)

(2107)

produces a translation of the wave function by a distance n a. ˆ φ(x − na) = S(na) φ(x)   ina = exp − pˆ φ(x) ¯h 443

(2108)

since the operator generates a Taylor-MacLaurin expansion in the displacement n a. Furthermore, a displacement of (n + m) a can be generated by successive displacements of n a and m a. The unitary transformation corresponding to the combined displacement can be expressed as ˆ ˆ ˆ ˆ ˆ S((n + m)a) = S(na) S(ma) = S(ma) S(na)

(2109)

Thus, the unitary operators corresponding to translations through any integer multiple of the basic periodicity a commute. Since under the displacement the potential transforms as ˆ V (x − na) = S(na) V (x) Sˆ† (na)

(2110)

and as the potential is periodic V (x − na) = V (x)

(2111)

ˆ S(na) V (x) Sˆ† (na) = V (x)

(2112)

one finds that ˆ and so S(na) commutes with the Hamiltonian, as ˆ = S(na) ˆ ˆ Sˆ† (na) H H

(2113)

ˆ or, on post multiplying by S(na), one has ˆ S(na) ˆ ˆ ˆ H = S(na) H

(2114)

ˆ and S(na), ˆ Thus, the set of operators H, form a set of mutually commuting operators and can be diagonalized. We seek a simultaneous solution of the set of eigenvalue equations ˆ φ(x) = E φ(x) H ˆ S(na) φ(x) = λn φ(x)

(2115)

for all n. The eigenvalues of a unitary operator must have modulus of unity, so one has | λn |2 = 1 (2116) Since, the discrete translation operator satisfies the equation ˆ ˆ ˆ S((n + m)a) = S(na) S(ma)

(2117)

the eigenvalues must be related by the multiplicative relation λn+m = λn λm

444

(2118)

Hence, the eigenvalues may be written as   λn = exp i n ϕ

(2119)

where ϕ is a real phase. This is a one-dimensional version of Bloch’s theorem, which states that eigenfunctions of a Hamiltonian with a periodic potential, with periodicity a, can be chosen such that   φ(x − na) = exp i n ϕ φ(x) (2120) ˆ Therefore, by demanding that φ(x) is an eigenfunction of S(na) we have ensured that the solution is finite at x → ± ∞. —————————————————————————————————— Example As an example, consider a periodic potential composed of repulsive delta functions of strength V0 a, V (x) =

n=+∞ X

V0 a δ( x − n a )

(2121)

n=−∞

Classically, a particle of energy E > 0 would be free to move in the regions between the repulsive delta functions. Quantum mechanically, we expect the particle to be able to tunnel through the narrow potential barrier. —————————————————————————————————— Solution The differential equation can be solved in the region between the barriers, say, in the region n a < x < ( n + 1 ) a the eigenfunction has the form     φ(x) = An exp i k x + Bn exp − i k x (2122) corresponding to a linear superposition of a forward and backward travelling waves. The wave vector k is related to the energy through E =

¯ 2 k2 h 2m

(2123)

The wave function in the region ( n + 1 ) a < x < ( n + 2 ) a also has the same form     φ(x) = An+1 exp i k x + Bn+1 exp − i k x (2124) 445

V(x) 20

15

10

5

0 -4

-3

-2

-1

0

1

2

3

4

x/a

Figure 110: A periodic array of repulsive delta function potentials. The pairs of coefficients (An , Bn ) and (An+1 , Bn+1 ) on each side of the delta function barrier at x = ( n + 1 ) a are related by demanding that the forms also satisfy the eigenvalue equation at the boundary. Continuity of the wave function at x = ( n + 1 ) a yields the equation     ( An+1 − An ) exp i k ( n + 1 ) a + ( Bn+1 − Bn ) exp − i k ( n + 1 ) a = 0 (2125) The discontinuity in the derivative at the boundary is related to the strength of the delta function barrier through     ( An+1 − An ) exp i k ( n + 1 ) a − ( Bn+1 − Bn ) exp − i k ( n + 1 ) a       2 m V0 a = An exp i k ( n + 1 ) a + Bn exp − i k ( n + 1 ) a i k ¯h2 (2126) This pair of equations can be solved to yield (An+1 , Bn+1 ) in terms of (An , Bn ). Thus, on adding the pair of equations one has   2 ( An+1 − An ) exp i k ( n + 1 ) a       2 m V0 a An exp i k ( n + 1 ) a + Bn exp − i k ( n + 1 ) a = ik¯ h2 446

(2127) while on subtracting them one obtains   2 ( Bn+1 − Bn ) exp − i k ( n + 1 ) a       2 m V0 a = − A exp i k ( n + 1 ) a + B exp − i k ( n + 1 ) a n n ik¯ h2 (2128) The pair of equations may be re-written as       m V0 a m V0 a An+1 = 1 + A + exp − 2 i k ( n + 1 ) a Bn n ik¯ h2 i k ¯h2 (2129) and       m V0 a m V0 a Bn+1 = 1 − Bn − exp + 2 i k ( n + 1 ) a An i k ¯h2 i k ¯h2 (2130) These two equations are not sufficient to determine a solution uniquely (up to an undetermined phase factor) since the ratio of the coefficients An and Bn are unknown. However, Bloch’s theorem gives the additional relation   φ(x − a) = exp i ϕ φ(x) (2131) which uniquely specifies the simultaneous eigenfunctions. On using Bloch’s theorem, one finds   φ(x − a) = exp i ϕ φ(x)       = exp i ϕ An+1 exp i k x + Bn+1 exp − i k x     = An exp i k ( x − a ) + Bn exp − i k ( x − a ) (2132) which, on equating the coefficients of the independent exponential functions, yields   An+1 = exp − i ( ϕ + k a ) An (2133) and

 Bn+1 = exp

 − i(ϕ − ka)

447

Bn

(2134)

This insures that the wave functions do not diverge in the limits x → ± ∞. On using these relations to eliminate Bn+1 and An+1 , one finds the two equations      m V0 a Bn exp − 2 i k ( n + 1 ) a An ik¯ h2     m V0 a = exp − i ( k a + ϕ ) − 1 + (2135) i k ¯h2     m V0 a An exp + 2 i k ( n + 1 ) a Bn ik¯ h2 "   # m V0 a = 1 − − exp + i ( k a − ϕ ) i k ¯h2 

(2136)

The above two equations can be combined to eliminate the ratio of An / Bn , which leads to the equation "    #"   # m V0 a m V0 a exp − i (ka + ϕ) − 1 + 1 − − exp + i (ka − ϕ) i k ¯h2 i k ¯h2  2 m V0 a = (2137) ik¯ h2 On re-arranging this equation, one finds      m V0 a + 1 = 0 exp − i 2 ϕ − 2 exp − i ϕ cos k a + sin k a k ¯h2 (2138)   Multiplying this equation by exp + i ϕ leads to  cos ϕ =

m V0 a cos k a + sin k a k ¯h2

! = 1

(2139)

This equation only has real solutions for ϕ, when the right hand side has a magnitude less than unity. In these ranges of k, the above equation can be solved to yield k as a function of ϕ. We note that ϕ is only defined up to multiples of 2 π. We shall examine the equation graphically, the right-hand side is plotted as a function of k a in fig(111). As a function of k, the function !  m V0 a (2140) F (k) = cos k a + sin k a k ¯h2 has a maximum at k = 0 with the value F (0) = 1 + 448

m V0 a2 ¯h2

(2141)

and behaves as cos ka in the asymptotic large k limit. The allowed values of k are those for which cos ϕ exists and, therefore, the function F (k) has a magnitude less than unity. Hence, not all values of k will give rise to a solution. In particular, the values of k for which | F (k) | > 1 do not yield a solution. The range of forbidden k values (at large k) are near k a = µ π, for integer µ.

6 5 4

F(k)

3 2 1 0 0

1

2

3

4

5

6

-1 -2

ka/π

Figure 111: The graphical determination of the allowed ranges of k. Since, E and k are related via E =

¯ 2 k2 h 2m

(2142)

and not all k values are allowed, we find that the spectrum of energy eigenvalues have “gaps” centered near the energies E =

¯ 2 µ2 π 2 h 2 m a2

(2143)

The allowed energies form bands, which are separated by band gaps. As k → ∞ the band gaps become very narrow. Since the discontinuities become negligible at high energies, the band energies closely follow the parabolic dispersion relation. Since k is not a good quantum number and due to the gaps in the E(k) relation, the energies are not plotted as E(k). However, ϕ is a good quantum ˆ number as it determines the eigenvalue of the S(a), thus the bands are usually plotted as E(ϕ), against ϕ. Typical energy bands are shown in fig(112). It is 449

seen that there are energy gaps between the bands at the origin ϕ = 0 and at the boundary ϕ = π. The sizes of the gaps diminish and the widths of the bands increase as one successively inspects higher energy intervals.

Dispersion Relation 36 30

E(φ φ)

24 18 12 6 0 0

0.25

0.5

0.75

1

φ/π

Figure 112: The band dispersion relation E(ϕ) plotted as a function of the Bloch eigenvalue ϕ. ——————————————————————————————————

4.9.4

Exercise 111

Calculate the energy eigenvalues associated with a potential constructed from a periodic array of attractive delta function potentials V (x) = −

n=∞ X

V0 a δ( x − n a )

(2144)

n=−∞

Explicitly, consider the eigenfunctions corresponding to E > 0 and E < 0. ——————————————————————————————————

450

4.9.5

Solution 111

The solution for the states with energy E > 0 and wave vector k can be expressed as     φ(x) = An exp i k x + Bn exp − i k x (2145) in the intervals between the delta functions. The relation between the coefficients An and Bn in successive intervals can be expressed in terms of the Bloch label ϕ. The Bloch label is given by the solution of the equation !  m V0 a ≤ 1 (2146) cos ϕ = cos k a − sin k a k ¯h2 Not all values of k yield solutions for ϕ, and as the energy eigenvalues are given by ¯h2 k 2 E = (2147) 2m one finds band gaps, for E > 0.

2

1

0

F(k)

0

1

2

3

4

5

6

-1

-2

-3

-4

ka/π

Figure 113: The graphical determination of the allowed ranges of k.

as

In addition to the E > 0 solutions, one has solutions which can be written     φ(x) = An exp + κ x + Bn exp − κ x (2148) 451

in each interval. These solutions correspond to negative energy eigenvalues ¯ 2 κ2 h (2149) 2m The eigenvalue of the translation operator can be expressed in terms of ϕ where !  m V0 a cos ϕ = cosh κ a − sinh κ a ≤ 1 (2150) κ ¯h2 E = −

This only produces solutions for one finite range of κ values. Hence, we find a single band with negative energies. The dispersion relation is shown in fig(115).

4 3 2

F(κ κ)

1 0 0

0.5

1

1.5

2

2.5

3

-1 -2 -3 -4

κa/π

Figure 114: The graphical determination of the allowed ranges of κ. ——————————————————————————————————

4.9.6

Rotational Invariance

A rotation of a system (through an angle θ about an arbitrary oriented axis represented by the unit vector eˆ) can be expressed in terms of a unitary operation    θ ˆ ˆ Ue (θ) = exp − i eˆ . J (2151) ¯h where Jˆ is the total angular momentum operator ˆ + S ˆ Jˆ = L 452

(2152)

35 30 25

E(φ φ)

20 15 10 5 0 -5 0

0.25

0.5

0.75

1

φ/π

Figure 115: The band dispersion relation E(ϕ) for the periodic array of attractive delta function potentials, plotted as a function of the Bloch eigenvalue ϕ. composed of the sum of the orbital angular momentum and the spin angular momentum operators. ˆe (θ) affects a rotation can be seen by its effect on the wave function That U of a spinless particle Ψ(r). Consider a rotation about the z axis through an infinitesimal angle δϕ = ϕN0 then Ψ0 (r)

ˆz (δϕ) Ψ(r) = U   δϕ ˆ = exp − i Lz Ψ( z eˆz + eˆy y + eˆx x ) ¯h   ∂ ∂ = 1 − δϕ ( x − y ) + O(δϕ2 ) Ψ(r) ∂y ∂x =

Ψ(r) + δϕ ( x eˆy − y eˆx ) . ∇ Ψ(r) + O(δϕ2 ) (2153)

and to lowest order in δϕ, we recognize this as a Taylor expansion of = Ψ(r + δϕ x eˆy − δϕ y eˆx ) When the position vector is represented as x = r sin θ cos ϕ 453

(2154)

Rotation of a scalar Ψ

Ψ(R-1r)

Ψ(r)

y-axis

r'=R-1r R

x-axis

ˆz (θ) produces a rotation of the scalar wave Figure 116: The unitary operator U ˆ acts function through an angle of θ around the z-axis. The rotation operator R ˆ r. on the coordinates and transforms the point r to the point r0 , i.e. r0 = R Under the rotation, the wave function Ψ(r) is transformed to Ψ0 (r), such that the value of Ψ0 at the point r0 is the same as the value of Ψ at the point r. Therefore, ˆ −1 r). The unitary operator which transforms Ψ0 (r0 ) = Ψ(r) so, Ψ0 (r) = Ψ(R ˆ (θ) Ψ(r) = Ψ(R ˆ −1 r). the wave function is then determined by Ψ0 (r) = U y z

= r sin θ sin ϕ = r cos θ

(2155)

we find that the transformed wave function is Ψ0 (r)

ˆz (δϕ) Ψ(r) = U = Ψ( z eˆz + r sin θ (sin ϕ − cos ϕδϕ) eˆy + r sin θ (cos ϕ + sin ϕδϕ) eˆx ) = Ψ( z eˆz + r sin θ sin(ϕ − δϕ) eˆy + r sin θ cos(ϕ − δϕ) eˆx ) (2156)

which has suffered an infinitesimal rotation of δϕ about the z axis. Thus, we find that the angular momentum operator is the generator for infinitesimal rotations. A finite rotation through ϕ0 can be built up from N successive infinitesimal transformations through δϕ = ϕ0 /N by taking the limit N → ∞ via ˆz (ϕ0 ) U

= =

lim

N → ∞

lim

N → ∞

ˆzN (δϕ) U  N ϕ0 ˆ 1 − i Lz N ¯h

454

 =

exp

− i

ϕ0 ˆ Lz ¯h

 (2157)

This could have been immediately recognized had we used the representation of ˆ z and the wave function in spherical polar coordinates both the operator L Ψ0 (r, θ, ϕ)

ˆz (ϕ0 ) Ψ(r, θ, ϕ) = U   ϕ0 ˆ = exp − i Lz Ψ(r, θ, ϕ) ¯h   ∂ = exp − ϕ0 Ψ(r, θ, ϕ) ∂ϕ = Ψ(r, θ, ϕ − ϕ0 )

(2158)

Rotation of a vector through ϕ

e exr ϕ

ex(exr) r(e.r) r

Figure 117: The rotation of an arbitrary vector r about an angle ϕ about the axis eˆ. An alternate derivation of the effect of a finite rotation through an angle ϕ0 about an arbitrary unit vector eˆ can be expressed in terms of its effect on the position vector. By composition, one can build up the effect of a rotation on an arbitrary wave function. Under a finite rotation, an arbitrary vector r has simple transformational properties as it behaves like a state with angular

455

momentum l = 1. After the rotation, the vector is given by40   r0 = ( r . eˆ ) eˆ − sin ϕ0 eˆ ∧ r + cos ϕ0 eˆ ∧ r ∧ eˆ   = ( r . eˆ ) eˆ − sin ϕ0 eˆ ∧ r + cos ϕ0 r − ( eˆ . r ) eˆ   = r − sin ϕ0 eˆ ∧ r + ( cos ϕ0 − 1 ) r − ( eˆ . r ) eˆ (2159) We shall use the identity  eˆ ∧ r =

 eˆ ∧ ( r ∧ ∇ )

r

(2160)

and a second identity  eˆ ∧ ( r ∧ eˆ ) = −

ˆ ) eˆ . ( r ∧ ∇

2 r

(2161)

Using these two identities, the rotated vector can be expressed as    2 0 r = r − sin ϕ0 eˆ . ( r ∧ ∇ ) r + ( 1 − cos ϕ0 ) eˆ . ( r ∧ ∇ ) r 

 1 − sin ϕ0

= 

 1 − i sin ϕ0

=  =

 eˆ . ( r ∧ ∇ )

1 +

ˆ eˆ . L ¯h

2 

 + ( 1 − cos ϕ0 )



 − ( 1 − cos ϕ0 )

ˆ eˆ . L ¯h

eˆ . ( r ∧ ∇ )

r

2  r

  ∞ ∞ X X ˆ  ˆ 2  ( − i ϕ0 )(2n+1) eˆ . L ( − i ϕ0 )(2n) eˆ . L + r (2n + 1)! ¯ h (2n)! ¯ h n=0 n=1 (2162)

where we have introduced the angular momentum operator. Since the component (ˆ e . r) of r is an eigenstate of ( eˆ . L ) with eigenvalue 0, and since the perpendicular components are eigenstates of ( eˆ . L )2 with eigenvalues + ¯h2 , we have   ˆ  ˆ 3 eˆ . L eˆ . L (2163) r = r ¯h ¯h 40 This

can be seen by introducing a basis consisting of three orthogonal unit vectors eˆ

1 eˆ ∧ rˆ sin θ 1 eˆθ = eˆ ∧ ( eˆ ∧ rˆ ) sin θ where cos θ = ( eˆ . rˆ ). In this basis, the vector r has components (cos θ, 0, − sin θ). The rotation through ϕ0 can then be performed in this coordinate system. eˆφ

=

456

This identity can be used to insert extra factors of  ˆ 2n eˆ . L ¯h

(2164)

ˆ into the general terms of both the even and odd series so that the powers of L match the powers of ϕ0 . The even and odd terms can be combined to form one series which exponentiates  X  ∞ ˆ n  ( − i ϕ0 )n eˆ . L 0 r = r n! ¯h n=0   ϕ0 = exp − i eˆ . L r (2165) ¯h giving the desired result. Rotations R can be combined to give other rotations. The rotations form a group. If a rotation is denoted by its axis eˆ and the angle of rotation θ, the rotation Re (θ) can be represented on a unit sphere. The rotation axes can be labeled by their points of intersection with the unit sphere. The combination of two rotations RP1 (θ1 ) and RP2 (θ2 ) can be found from Euler’s construction shown in figs (118) and (119), and is neatly expressed algebraically in terms of multiplication of quaternions. Euler’s construction shows that rotations do not commute if the rotations do not share the same axis, and so groups of rotations can be non-Abelian. We note that if the Hamiltonian of a system is invariant under rotations, then the angular momentum operator commutes with the Hamiltonian. If this is the case, then angular momentum is conserved. For a particle with spin, the rotation operator consists of a product of two factors. The first factor acts on the spatial components and is identical to the rotation operator for the spinless particle. The second factor acts (locally) on the components of the spinor. For a spin one half particle, the part of the unitary operator that acts on the spin which represents a rotation through ϕ0 about the axis eˆ is given by   ˆe (ϕ0 ) = exp − ϕ0 i eˆ . σ U 2 ϕ0 ϕ0 = cos (2166) σ0 − i sin eˆ . σ 2 2 This unitary operator41 acts on the two-component spinor wave function. We should note that a rotation by 2 π does not leave the spinor invariant but changes 41 An arbitrary spin one half rotation operator can be represented by unitary 2 × 2 matrices, with determinant +1. Hence, the matrices are the special unitary matrices of SU (2).

457

Euler Construction

P1 θ1/2

θ1/2

P4

θ3/2 θ2/2 θ2/2

P2

P3

π−θ3/2

R1

O

Figure 118: The Euler construction. The rotation RP1 (θ1 ) through an angle θ1 about the axis OP1 is to be combined with the rotation RP2 (θ2 ) through an angle θ2 about the axis OP2 . Two auxiliary great circles are constructed from P1 which subtend angles θ21 to the great circle connecting P1 and P2 . Likewise, two more auxiliary great circles are constructed emanating from point P2 . The points of intersection of the pairs of auxiliary great circles are labeled as P3 and P4 . It is seen that RP1 (θ1 ) shifts P3 to position P4 , and is shifted back to P3 by the subsequent application of RP2 (θ2 ). Hence, P3 is on the axis of the combined rotation. Also P4 is on the axis of a combined rotation when the component rotations are combined in the opposite order. its phase by π. Thus, a spin half particle should be rotated through 4 π before the initial state is recovered42 . Thus, for example, a rotation of a spin by ϕ0 about the z-axis is produced by ˆz (ϕ0 ) = U



exp[ − i 0

ϕ0 2

]

0 exp[ + i

 ϕ0 2

]

(2167)

That this operator produces a rotation of the spin state can be verified by letting it act on the eigenstate of the operator η . Sˆ with eigenvalue + h¯2 . In this 42 In a recent experiment, spin one half particles were sent through a two-channel interferometer. A magnetic field was used to rotate the particles in one channel. The interference pattern was periodic in the rotation angle, with period 4 π. The experiment was described in A.W. Overhauser, A.R. Collela and S.A. Werner, Phys. Rev. Lett. 33, 1237 (1974).

458

Euler Construction P2 θ2/2 θ2/2

P1

θ1/2

θ3/2

π−θ3/2

P3

R

O

Figure 119: The Euler construction. The angle of the combined rotation can be found by considering the effect of the successive rotations on point P1 . Under the first transformation P1 is invariant, since it is on the axis of rotation. Under the second transformation P1 is swept to point R. Thus, under the combined rotation around P3 , the great circle segment P3 P1 is swept to P3 R. Since P2 P3 bisects the segments P2 P1 and P2 R, the angle of the combined rotation is found to be 2(π − θ23 ). expression, η is a unit vector that has the Cartesian components   sin θ cos ϕ η =  sin θ sin ϕ  cos θ The η component of the spin operator is given by   h ¯ cos θ sin θ exp[ − i ϕ ] η . Sˆ = sin θ exp[ + i ϕ ] − cos θ 2 and the eigenstate with eigenvalue + h¯2 is given by   cos θ2 exp[ − i ϕ2 ] Ψη = ϕ sin θ2 exp[ + i 2 ]

(2168)

(2169)

(2170)

This eigenstate describes a spin pointing along the positive η direction. The rotated spin state is then given by Ψ0η0

ˆz (ϕ0 ) Ψη = U 459

 =

0 cos θ2 exp[ − i ϕ+ϕ ] 2 ϕ+ϕ0 θ sin 2 exp[ + i 2 ]

 (2171)

Hence, we deduce that Ψ0η0 is the eigenstate of η 0 . S with eigenvalue + h¯2 , where the unit vector η 0 is obtained by rotating η around the z axis through the angle ϕ0 . That is, the spins now point along the positive η 0 direction. The operator ˆz (ϕ0 ) produced a rotation of the spin about the z-axis through an angle of ϕ0 . U The combined rotation on the spin and spatial wave function of the spin one half particle   Ψ+ (r) Ψ(r) = (2172) Ψ− (r) produces the transformed state  0  0 Ψ+ (r) Ψ (r) = 0 Ψ− (r)     Ψ+ (Rz−1 r) exp[+i ϕ20 ] 0 = Ψ− (Rz−1 r) 0 exp[−i ϕ20 ]

(2173)

In this expression, the spinor wave function is transferred from the point Rz−1 r to the rotated point r and the direction of the spin is adjusted locally. For a particle with spin one, the corresponding (local) spin rotation operator can be represented as   i ϕ0 ˆ Ue (ϕ0 ) = exp − eˆ . S ¯h     2 eˆ . S eˆ . S = Iˆ − i sin ϕ0 + cos ϕ0 − 1 ¯h ¯h (2174) For integer spins, a rotation of 2 π does not produce a change in the phase of spinor wave function. This is in contrast to the case of half integer spins where the phase of the wave function changes by π. As an example, a spin is rotated through an angle ϕ0 about the z-axis by ˆz (ϕ0 ) where the transformation U   exp[ − i ϕ0 ] 0 0 ˆz (ϕ0 ) =   0 1 0 U (2175) 0 0 exp[ + i ϕ0 ] The state in which the direction of the classical spin is known to be along an arbitrary direction η corresponds to the quantum state which is an eigenstate

460

of the component of the spin along η is h ¯ . The relevant operator is given by   √1 sin θ exp[ − i ϕ ] cos θ 0   2 η.S  √1 sin θ exp[ − i ϕ ]  0 =  √12 sin θ exp[ + i ϕ ]  2 h ¯ √1 sin θ exp[ + i ϕ ] 0 − cos θ 2 (2176) where θ and ϕ are the polar coordinates that specify the direction of the unit vector η. The operator has an eigenstate with eigenvalue ¯h which is given by   ( 1 + cos √θ ) exp[ − i ϕ ] 1  Ψη = (2177) 2 sin θ 2 ( 1 − cos θ ) exp[ + i ϕ ] Hence, this eigenstate describes a spin pointing in the direction of positive η. This result is consistent with the calculated expectation values of Sˆx , Sˆy and ˆz (ϕ0 ) results in a Sˆz for this state. It is simple to see that the transformation U ˆ spin state Uz (ϕ0 ) Ψη where the spin’s direction is rotated by ϕ0 . ——————————————————————————————————

4.9.7

Exercise 112

ˆe (ϕ0 ), for a spin one half particle, satisfies Show that the spin rotation operator U the identity   ˆe (ϕ0 ) = exp − ϕ0 i eˆ . σ U 2 ϕ0 ϕ0 = cos σ0 − i sin eˆ . σ (2178) 2 2 ——————————————————————————————————

4.9.8

Solution 112

First we expand the operator, and rearrange the series as a sum of even and odd terms   ϕ0 ˆ Ue (ϕ0 ) = exp − i eˆ . σ 2  n ∞ X 1 ϕ0 = − i eˆ . σ n! 2 n=0 (2n+1) 2n   ∞ ∞ X X 1 ϕ0 1 ϕ0 + − i eˆ . σ = − i eˆ . σ (2n + 1)! 2 2n! 2 n=0 n=0 (2179) 461

Then prove the identity 2

 eˆ . σ

= σ0

(2180)

which proceeds from the expansion of the scalar product  2  X 2 eˆ . σ = eˆ . eˆi σi i

=

X

( eˆ . eˆi ) ( eˆ . eˆi ) σi σj

i,j

=

X

( eˆ . eˆi )2 σi2 +

X

i

i>j

X

( eˆ . eˆi )2 σ0 +

X

=

i

 ( eˆ . eˆi ) ( eˆ . eˆi )

 σi σj + σj σi

( eˆ . eˆi ) ( eˆ . eˆi ) δi,j

i>j

= ( eˆ . eˆ ) σ0 = σ0

(2181)

since the squares of the Pauli matrices are unity and the different Pauli matrices anti-commute. The last line follows as eˆ is a unit vector. On substituting the identity in the series expansion, one finds   ˆe (ϕ0 ) = exp − i ϕ0 eˆ . σ U 2  2n  2n+1 ∞ ∞ n X X ϕ0 ( − 1 )n ϕ0 (−1) σ0 − i ( eˆ . σ ) = 2n! 2 (2n + 1)! 2 n=0 n=0 (2182) Since, the trigonometric functions are defined as  2n ∞ X ϕ0 ( − 1)n ϕ0 cos = 2 (2n)! 2 n=0  (2n+1) ∞ X ϕ0 ( − 1)n ϕ0 sin = 2 (2n + 1)! 2 n=0

(2183)

one has proved the identity Uˆe (ϕ0 ) = =



 i exp − ϕ0 eˆ . σ 2 ϕ0 ϕ0 cos σ0 − i sin eˆ . σ 2 2

(2184)

——————————————————————————————————

462

4.9.9

Exercise 113

Consider the rotation of an arbitrary (spatially uniform) spin one half state Ψ. ˆz (ϕ0 ). Express the The rotation is represented by the unitary transformation U expectation value of the x-component of spin in the transformed state Ψ0 in terms of the untransformed states. Hence, show that 0 0 ( Ψ † Sˆx Ψ ) = cos ϕ0 ( Ψ† Sˆx Ψ ) − sin ϕ0 ( Ψ† Sˆy Ψ )

(2185)

——————————————————————————————————

4.9.10

Solution 113 0

The transformed state is given by Ψ where 0 ˆz (ϕ0 ) Ψ Ψ = U

(2186)

Hence, the expectation value is given by 0 0 ˆ † (ϕ0 ) Sˆx U ˆz (ϕ0 ) Ψ Ψ † Sˆx Ψ = Ψ† U z

(2187)

In matrix form, we have       0 0 h † ¯ 0 exp[+i ϕ20 ] 0 0 1 exp[−i ϕ20 ] † ˆ Ψ Ψ Sx Ψ = Ψ 0 exp[+i ϕ20 ] 0 exp[−i ϕ20 ] 1 0 2   0 exp[+iϕ0 ] = Ψ† Ψ exp[−iϕ0 ] 0 = cos ϕ0 ( Ψ† Sˆx Ψ ) − sin ϕ0 ( Ψ† Sˆy Ψ ) (2188) The expectation value of Sˆy also changes like the y-component of a classical vector under rotation, as does the expectation value of Sˆz . The expectation value of Sˆz doesn’t change since the rotation is about the z-axis. Hence, the expectation value of Sˆ of any arbitrary state transforms like a vector. —————————————————————————————————— The Vector Character of Spin One Particles For a particle with spin one, the corresponding (local) spin rotation operator can be represented as   i ϕ0 ˆ Ue (ϕ0 ) = exp − eˆ . S ¯h     2 eˆ . Sˆ eˆ . Sˆ = Iˆ − i sin ϕ0 + cos ϕ0 − 1 ¯h ¯h (2189) 463

ˆz (ϕ0 ) which produces a rotation of the spin of Therefore, the transformation U a spin one particle through the angle ϕ0 about the z-axis is given by   exp[ − i ϕ0 ] 0 0 ˆz (ϕ0 ) =   0 1 0 (2190) U 0 0 exp[ + i ϕ0 ] in the basis formed by the set of eigenstates of Sˆz . Likewise, the forms of the ˆy and U ˆx that produce spin rotations about the y and x axes are operators U given by  1  1 √1 2 ( 1 − cos ϕ0 ) 2 ( 1 + cos ϕ0 ) − 2 sin ϕ0  ˆy (ϕ0 ) =  √1 sin ϕ0 cos ϕ0 − √12 sin ϕ0 U   (2191) 2 1 1 √1 sin ϕ0 ( 1 − cos ϕ ) ( 1 + cos ϕ ) 0 0 2 2 2 and 1 + cos ϕ0 ) − √i2 sin ϕ0  ˆx (ϕ0 ) =  − √i2 sin ϕ0 cos ϕ0 U 1 i √ 2 ( cos ϕ0 − 1 ) − 2 sin ϕ0 

1 2(

1 2(

 cos ϕ0 − 1 )  − √i2 sin ϕ0  1 ( 1 + cos ϕ ) 0 2

(2192)

ˆ given by The operator Q  −1 0 1 1 √  − i √0 − i  2 0 2 0 

ˆ = Q

(2193)

represents a unitary transformation. This operator transforms states from the e via basis formed by the eigenstates of Sˆz (Ψ) to a new basis (Ψ)     ex Ψ Ψ+1  e  ˆ  Ψ0  (2194)  Ψy  = Q e Ψ −1 Ψz The z-component of the transformed spin operator is given by e Sˆz

ˆ Sˆz Q ˆ† = Q  −1 0 h  ¯ − i √0 = 2 0 2  0 −i 0 0 0 = h ¯  i 0 0 0

  1 1 −i   0 0 0 

  −1 0 0 0 0   0 0 −1 1

 i √0 0 2  i 0 (2195)



On applying the unitary transform to the  0 h ¯ Sˆy = √  i 2 0 464

y-component of the spin operator  −i 0 0 −i  (2196) i 0

one obtains the transformed operator as e Sˆy

ˆ Sˆy Q ˆ† = Q  −1 h  ¯ −i = 2 0  0 = h ¯  0 −i

  0 0 1 i √   0 − i  2 √ 2 0 0  0 i 0 0  0 0



√i 2

0 √i 2



0 √i 2

0

  −1   0  1

 i √0 0 2  i 0 (2197)

Likewise, applying the transformation to the x-component   0 1 0 ¯ h Sˆx = √  1 0 1  2 0 1 0

(2198)

one obtains the transformed operator e Sˆx

ˆ Sˆx Q ˆ† = Q    0 −1 0 1 h  ¯  √1  − i √0 − i =  2 2 0 2 0 0   0 0 0 = h ¯  0 0 −i  0 i 0

√1 2

0 √1 2

0 √1 2

0

  −1 i   0 0  1 i

 √0 2  0 (2199)

e Therefore, the (j, k) matrix element of the transformed spin operator Sˆi is simply given by e  Sˆi j,k = − i ¯h εi,j,k (2200) where εi,j,k is the Levi-Civita symbol. Since the transformed and untransformed operators are related via a unitary transformation, the transformed operators obey the same commutation relations as the untransformed operators. In particular, the squared magnitude of the transformed spin operator f Sˆ2

f f f = Sˆx2 + Sˆy2 + Sˆz2   1 0 0 2  0 1 0  = 2h ¯ 0 0 1

(2201)

is diagonal and, therefore, commutes with all other spin operators. e ˆ e (ϕ0 ) in the new basis is given by The form of the spin rotation operator U e ˆ e (ϕ0 ) = Q ˆU ˆe (ϕ0 ) Q ˆ† U 465

(2202)

where e ˆ e (ϕ0 ) U



 i ϕ0 eˆ = exp − eˆ . S ¯h    e 2 e  eˆ . Sˆ eˆ . Sˆ ˆ + cos ϕ0 − 1 = I − i sin ϕ0 ¯h ¯ h (2203)

In particular, the matrix which describes a rotation about the z-axis is given by   cos ϕ0 − sin ϕ0 0 e ˆ z (ϕ0 ) =  sin ϕ0 cos ϕ0 0  U (2204) 0 0 1 The matrix which describes a rotation about  cos ϕ0 e ˆ y (ϕ0 ) =  0 U − sin ϕ0

the y-axis becomes  0 sin ϕ0  1 0 0 cos ϕ0

and a rotation about the x-axis is produced by   1 0 0 e ˆ x (ϕ0 ) =  0 cos ϕ0 − sin ϕ0  U 0 sin ϕ0 cos ϕ0

(2205)

(2206)

These transformations are recognized as being identical to the expressions for the transformations of vectors under rotations through a finite angle ϕ0 about the Cartesian axes. A more general (local) rotation of the spin one wave function can be built up from infinitesimal rotations about an arbitrary axis. Under an infinitesimal rotation δϕ0 about an arbitrary axis eˆ, the spin one wave function changes (locally) according to     e 0x ex     Ψ Ψ e ˆ  e   e0  ˆ1 − i δϕ0 eˆ . S (2207)  Ψy   Ψy  = h ¯ 0 e e Ψz Ψz In terms of the components, the transformation is expressed as  f  X  X eˆk . Sˆk e 0i = ej Ψ δi,j − i δϕ0 Ψ ¯h i,j j k  X  X ej = δi,j − δϕ0 eˆk εk,i,j Ψ j

(2208)

k

If one interprets the components of wave function of the spin one particle as a vector in the same space as ordinary vectors, then on re-labelling and rearranging the indices, one finds X X ej = ek − eˆk εk,i,j Ψ eˆj εi,j,k Ψ j,k

j,k

466

Rotation of a Vector Field Ψ'(r)=exp[-iSzφ] Ψ(Rz-1r)

y

Ψ(r)

Rz

x

ˆz (ϕ) produces a rotation of the spin one Figure 120: The unitary operator U wave function through an angle of ϕ around the z-axis. The rotation operator ˆ z acts on the coordinates and transforms the point r to the point r0 , i.e. R ˆ z r. Under the rotation, the vector wave function Ψ(r) is transformed to r0 = R Ψ0 (r), such that the value of Ψ0 at the point r0 is the same as the value of Ψ at the point r. However, the rotation also rotates the direction of the vector field ˆ −1 r). by ϕ. Therefore, Ψ0 (r0 ) = exp[−iϕ Sˆz ] Ψ(r) so, Ψ0 (r) = exp[−iϕ Sˆz ] Ψ(R The unitary operator which transforms the wave function is then determined ˆ (ϕ) Ψ(r) = exp[−iϕ Sˆz ] Ψ(R ˆ −1 r). by Ψ0 (r) = U  =

 e eˆ ∧ Ψ

(2209) i

Therefore, the spin one wave function transforms under rotations via    e + δϕ0 eˆ ∧ Ψ e e0 = Ψ Ψ

(2210)

This is the same transformation law obeyed by a vector under a rotation through an infinitesimal angle δϕ0 about the axis eˆ. This suggests that the wave function of a spin one particle should be considered as a vector field. If the spin one wave function Ψ has a vector representation, then its decomposition in terms of Cartesian unit vectors eˆi and the Cartesian components Ψi is given by (2211) Ψ = eˆx Ψx + eˆy Ψy + eˆz Ψz

467

Sometimes it may be convenient to express the vector field in terms of components Ψm which have definite values of Sˆz . The components of the vector in the two basis sets are related by Ψm

ˆ † Ψi = Q

or, equivalently, in matrix form by    −1 Ψ+1 1  Ψ0  = √  0 2 Ψ−1 1

   i √0 Ψx 2   Ψy  0 Ψz i 0

(2212)

(2213)

——————————————————————————————————

4.9.11

Exercise 114

f Find the spin states that are the simultaneous eigenstates of the operators Sˆz f and Sˆ2 . Show that, with an appropriate choice of phase, these states can be mapped e which are expressed in terms of the onto the circularly polarized states Ψ m Cartesian basis by   1 1 e  √ i  Ψ = − +1 2 0   0 e  0  Ψ = 0 1   1 1 e Ψ (2214) = √  −i  −1 2 0

——————————————————————————————————

4.9.12

Solution 114

On expressing the eigenvectors in the form  e Ψ m

 α =  β  γ 468

(2215)

one finds that the form of the eigenstates trivially satisfy the eigenvalue equation f f for Sˆ2 . The eigenvalue equation for the operator Sˆz is non-trivial and can be written as       0 −i 0 α α 0 0   β  = λm  β  h  i ¯ (2216) 0 0 0 γ γ The eigenvalues λm are determined from the zeros of the determinant −λ i 0

−i −λ 0

0 0 −λ

= 0

(2217)

which yield the solutions for λm as (1, 0, −1). The eigenvectors for λ = ± 1 are determined from the equations       0 −i 0 α±1 α±1 0 0   β±1  = ±  β±1  h  i ¯ (2218) 0 0 0 γ±1 γ±1 which yields the equations − i β±1 = i α±1 = 0 =

± α±1 ± β±1 ± γ±1

(2219)

Since these are homogeneous linear equations, the (un-normalized) solutions can be found by choosing α±1 = 1. The eigenvector corresponding to λm = 0 is given by − i β0 = i α0 = 0 =

0 0 0

(2220)

which can be solved by choosing γ0 = 1. On normalizing the above set of eigenstates, one finds that they can be expressed in forms   1 1 e  √ i  = Ψ +1 2 0   1 1 e Ψ = √  −i  −1 2 0   0 e  0  Ψ = (2221) 0 1 469

These correspond to the given set of eigenstates, except for the difference in e 1 . The desired phase could have been arrived at by the choice of phase for Ψ choosing α1 = − 1. —————————————————————————————————— Exercise 115 Consider a general rotation of a vector through an angle θ in d dimensions. ˆ (θ) is given by Show that trace of the rotation operator U ˆ (θ) = ( d − 2 ) + 2 cos θ Trace U

(2222)

—————————————————————————————————— When expressed in terms of the angular momentum basis, an arbitrary vector Ψ has the decomposition Ψ =

m=+1 X

eˆm Ψm

(2223)

m=−1

which defines the appropriate angular momentum basis vectors eˆm . The basis vectors eˆm are given in terms of the Cartesian basis vectors via eˆ+1 eˆ0 eˆ−1

1 = − √ ( eˆx + i eˆy ) 2 = eˆz 1 = √ ( eˆx − i eˆy ) 2

(2224)

These complex basis vectors are the circular polarization vectors for a spin one particle. The set of complex basis vectors eˆm for the angular momentum representation satisfy eˆ∗m = (−1)m eˆ−m

(2225)

and their scalar products are given by eˆ∗m . eˆm0 = (−1)m eˆ−m . eˆm0 = δm,m0

(2226)

The components of the vector Ψ would then be found from Ψm = eˆ∗m . Ψ

(2227)

A similar type of decomposition in terms of circularly polarized basis vectors is frequently applied to the vector wave function of the spin one photon.

470

4.9.13

Gauge Invariance

The Schr¨ odinger equation for a charged particle in an electromagnetic field is given by i ¯h

∂ Ψ(r, t) = ∂t



1 2m

 − i ¯h ∇ −

q A(r, t) c

2

 + q φ(r, t)

Ψ(r, t) (2228)

The vector and scalar potentials are invariant under the gauge transformations A(r, t) → A0 (r, t) = A(r, t) + ∇ Λ(r, t) 1 ∂ φ(r, t) → φ0 (r, t) = φ(r, t) − Λ(r, t) c ∂t

(2229)

The primed and unprimed potentials are in different gauges, but represent the same physical system. The Schr¨odinger equation in the primed gauge is ∂ 0 Ψ (r, t) i ¯h ∂t

 =

1 2m



q 0 − i ¯h ∇ − A (r, t) c

2

0

+ q φ (r, t)



Ψ0 (r, t) (2230)

where the primed wave function is given in terms of the unprimed wave function by   q 0 Ψ(r, t) → Ψ (r, t) = Ψ(r, t) exp i Λ(r, t) (2231) ¯h c and where the scalar field involved in the gauge transformation has been absorbed into the phase of the wave function. The gauge transformation is produced by the unitary transformation U such that Ψ0 (r, t) = U Ψ(r, t)   q U = exp i Λ(r, t) ¯h c

(2232)

Then the Hamiltonians in the two gauges are connected via ˆ0 = U H ˆ U † − i ¯h U ∂ U † H ∂t

(2233)

as can be seen by examining i ¯h

∂ ˆ Ψ(r, t) Ψ(r, t) = H ∂t

(2234)

Ψ(r, t) = U † Ψ0 (r, t)

(2235)

which on substituting

471

becomes U† i ¯ h



∂ 0 Ψ (r, t) + i ¯h ∂t

∂ † U ∂t



ˆ U † Ψ0 (r, t) Ψ0 (r, t) = H

(2236)

On utilizing the primed Schr¨odinger equation i¯ h

∂ 0 ˆ 0 Ψ0 (r, t) Ψ (r, t) = H ∂t

(2237)

one has ˆ 0 Ψ0 (r, t) + i ¯h U H †



∂ † U ∂t



ˆ U † Ψ0 (r, t) Ψ0 (r, t) = H

(2238)

This equation determines the time evolution of an arbitrary initial wave function Ψ0 (r, t), and so we have an operator equation   ∂ † † ˆ0 ˆ U† U H + i ¯h U = H (2239) ∂t ˆ and H ˆ 0. which on pre-multiplying by U is identical to the relation between H We note that this gauge transformation leaves r and the velocity pˆ − qc A(r, t) invariant. This can be seen as r0 = U r U † = r

(2240)

and pˆ0 −

q 0 A (r, t) c



 q A(r, t) U † c q q = pˆ − A(r, t) − ∇ Λ(r, t) c c q 0 = pˆ − A (r, t) c = U

pˆ −

(2241)

Finally, as p ˆ0 = pˆ = − i ¯ h ∇, the above quantity is gauge invariant as was to be proved. Thus, the Schr¨ odinger equation is gauge invariant. ——————————————————————————————————

4.9.14

Exercise 116

Find the continuity equation for a charged particle in an electromagnetic field, and show that the appropriate probability current density j(r, t) and the charge density ρ(r, t) are gauge invariant. ——————————————————————————————————

472

4.9.15

Solution 116

The equation of motion for the wave function Ψ is   2  ∂ 1 q i ¯h Ψ = pˆ − A + qφ Ψ ∂t 2m c or ∂ ih ¯ Ψ = ∂t



1 2m



q − i ¯h ∇ − A c

2

(2242)

 + qφ

Ψ

and the complex conjugate Ψ∗ satisfies 2    1 q ∂ ∗ − i ¯h Ψ = + i ¯h ∇ − A + q φ Ψ∗ ∂t 2m c

(2243)

(2244)

Multiplying the first equation by Ψ∗ and the second by Ψ, then subtracting the second from the first, one obtains     ∂ ¯h2 i ¯h Ψ∗ Ψ = − Ψ∗ ∇2 Ψ − Ψ ∇2 Ψ∗ ∂t 2m   q Ψ∗ A(r, t) Ψ (2245) + i ¯h ∇ mc Thus, one has the continuity equation ∂ρ + ∇j = 0 ∂t

(2246)

but ρ(r, t)

=

j(r, t)

=

Ψ∗ (r, t) Ψ(r, t)   h ¯ q ∗ ∗ =m Ψ ∇ Ψ − Ψ ∇ Ψ − Ψ∗ (r, t) A(r, t) Ψ(r, t) m mc (2247)

The charge density ρ(r, t) is gauge invariant, as under the gauge transformation one has   q 0 (2248) Λ(r, t) Ψ(r, t) Ψ(r, t) → Ψ (r, t) = exp i c ¯h and the phase factor drops out of the density. In the current density, the gradient factor transforms as   q Ψ∗ ∇ Ψ → Ψ0∗ ∇ Ψ0 = Ψ∗ ∇ Ψ + i Ψ∗ ∇ Λ Ψ c ¯h (2249) 473

and the term involving the vector potential transform as Ψ∗ A(r, t) Ψ → Ψ0∗ A0 (r, t) Ψ0

(2250)

where Ψ0∗ A0 (r, t) Ψ0

=

Ψ∗ A(r, t) Ψ + Ψ∗ ∇Λ(r, t) Ψ

(2251)

Thus, the factor in the current density involving ∇ Λ cancels, leaving j(r, t) gauge invariant. ——————————————————————————————————

4.9.16

Galilean Boosts

A transformation to a uniformly moving reference frame is represented by a time-dependent unitary transform. The (passive) transformation from a stationary reference frame to one moving with a velocity - u is represented by the ˆ defined by unitary matrix U   i ˆ U = exp − u . ( m r − pˆ t ) (2252) ¯h The above Galilean transformation is derived as the non-relativistic limit of a Lorentz boost, which is just a (static) rotation in space-time. However, in the ˆ is replaced by the rest mass energy mc2 . non-relativistic limit the operator H The above transformation represents a Galilean boost since it has the effect that r0 pˆ0

ˆ† = r + u t ˆ rU = U ˆ pˆ U ˆ † = pˆ + m u = U

(2253)

as expected for a transformation to a reference frame moving with constant velocity. The Hamiltonian is the generator of time-translations and, therefore, a state Ψ evolves according to the prescription   ∂ ∗ ˆ Ψ Ψ i ¯h Ψ = Ψ∗ H (2254) ∂t on transforming from Ψ to Ψ0 via ˆ Ψ Ψ0 = U one finds that   ∂ ∗0 ˆ H ˆ U ˆ † Ψ0 − i ¯h Ψ∗0 U ˆ ∂ U ˆ † Ψ0 Ψ0 = Ψ∗0 U Ψ i¯ h ∂t ∂t 474

(2255)

(2256)

ˆ 0 as which defines the transformed Hamiltonian H ˆ0 H

ˆ H ˆ U ˆ † − i ¯h U ˆ ∂ U ˆ† = U ∂t ˆ H ˆ U ˆ† − u . p = U ˆ

(2257)

Thus, the energy eigenvalues are related via E0 = E − u . p

(2258)

which is recognized as the non-relativistic limit of the Lorentz transformation of the time-like component of a four-vector E − u.p E0 = p 1 − ( uc )2 For a Hamiltonian in the stationary reference frame of the form  2 1 q ˆ H = pˆ − A(r) + V (r) 2m c

(2259)

(2260)

the Hamiltonian in the moving reference frame has the form  2 1 q ˆ0 = H pˆ + m u − A(r + ut) + q φ(r + ut) − u . pˆ 2m c  2 1 q q m u2 = pˆ − A(r + ut) + q φ(r + ut) − u . A(r + ut) + 2m c c 2 (2261) when expressed in terms of the untransformed position and momentum operators. In the above equations we have ignored any explicit time-dependence of the electromagnetic potentials. The transformed Hamiltonian is increased by an additive constant equal to the kinetic energy of the boost. Since we are working in the non-relativistic limit, this additive constant has no physical consequence. The vector potential is transformed according to A(r) → A0 (r0 ) = A(r + ut)

(2262)

which is the same as in the static reference frame (except that it is evaluated at the transformed position r0 = r + ut), but the scalar potential is transformed according to 1 φ(r) → φ0 (r0 ) = φ(r0 ) − u . A(r0 ) (2263) c This has the implication that the magnetic field in the moving frame defined by B 0 (r0 ) = ∇ ∧ A(r0 )

(2264)

is simply given by the magnetic field at the transformed position B 0 (r0 ) = B(r + ut) 475

(2265)

On the other hand, the transformed electric field defined by E 0 (r0 ) = − ∇ φ0 (r0 ) −

1 ∂ 0 0 A (r ) c ∂t

(2266)

is evaluated as   1 1 ∂ 0 E (r ) = − ∇ φ(r ) + ∇ u . A(r ) − A(r + ut) c c ∂t     1 1 1 ∂ 0 0 = − ∇ φ(r ) + ∇ u . A(r ) − u . ∇ A(r0 ) − A(r0 ) c c c ∂t (2267) 0

0

0

where the last term is a partial derivative with respect to t at constant r0 = r+ut. On using the identity       ∇ u.A − u.∇ A = u ∧ ∇ ∧ A (2268) one obtains the transformed electric field as 1 u ∧ B(r0 ) E 0 (r0 ) = E(r0 ) + c which is as might have been expected.

5

(2269)

The Rotating Planar Oscillator

To end this first semester of quantum mechanics, we shall look at a simple system and show that there are still some surprises to be found in separable systems. Usually systems are neither separable, nor can their excitations be found analytically. Most systems have to be solved numerically and it expected that their dynamics may sometimes show chaotic motion. We shall consider a two-dimensional simple harmonic oscillator, that is slowly being rotated about the origin with frequency Ω. The classical Lagrangian is given by   m m 2 2 L = r˙ 2 + r2 ( ϕ˙ − Ω )2 − ω r (2270) 2 2 The generalized momenta are given by pr pϕ

= m r˙ = mr

2



 ϕ˙ − Ω

(2271)

Therefore, the classical Hamiltonian can be expressed as H

= pr r˙ + pϕ ϕ˙ − L =

p2ϕ p2r m 2 2 + + pϕ Ω + ω r 2m 2 m r2 2 476

(2272)

In two dimensions, the angular momentum pϕ (which we shall now denote by L) is masquerading as a scalar. Hence, we shall write the Hamiltonian as H =

L2 m 2 2 p2r + + LΩ + ω r 2m 2 m r2 2

(2273)

We shall quantize this classical Hamiltonian and shall determine the energy eigenfunctions in the form   1 exp i l ϕ (2274) Ψnr ,l (r, ϕ) = Rnr ,l (r) √ 2π where l is a positive or negative integer. The radial part of the energy eigenvalue equation takes the form    2 2  h2 1 ∂ ¯ ∂R ¯h l m 2 2 − r + +h ¯lΩ+ ω r R = Enr ,l R (2275) 2 m r ∂r ∂r 2 m r2 2 We shall introduce a dimensionless variable ρ via the definition r mω ρ = r ¯h

(2276)

and re-write the radial equation as    2  1 ∂ ∂Rnr ,l l Ω 2 Enr ,l 2 ρ + + 2 l + ρ Rnr ,l = Rnr ,l (2277) ρ ∂ρ ∂ρ ρ2 ω ¯ ω h For Ω = 0, the radial equation can easily be solved by using the operator algebra of raising and lowering operators. The raising operators are expressed as     ∂ 1 2 l +1 Aˆ†l = − + + ρ − (2278) ∂ρ 2ρ 2ρ and the lowering operators are defined as the Hermitean conjugate operators43     1 2 l +1 ∂ ˆ Al = + + ρ − (2279) ∂ρ 2ρ 2ρ The commutator of the raising and lowering operator is evaluated as 2l + 1 [ Aˆ†l , Aˆl ] = − 2 − ρ2

(2280)

43 The operators contain terms that depend on the dimensionality. In d dimensions, the corresponding pair of Hermitean conjugate operators are given by

ˆ† = − A l



∂ d − 1 + ∂ρ 2ρ



 +

2l + d − 1 ρ − 2 ρ



and

 ˆl = A

∂ d − 1 + ∂ρ 2ρ



 +

477

2l + d − 1 ρ − 2 ρ



The product of the lowering and raising operators is evaluated as ∂2 1 ∂ l2 Aˆ†l Aˆl = − − + 2 + ρ2 − 2 ( l + 1 ) 2 ∂ρ ρ ∂ρ ρ

(2281)

which has a form similar to the left-hand side of the radial equation44 . On using the commutator in eqn(2280), one finds that the product of the lowering and raising operators taken in opposite order is given by ∂2 1 ∂ ( l + 1 )2 Aˆl Aˆ†l = − − + + ρ2 − 2 l 2 ∂ρ ρ ∂ρ ρ2

(2283)

The dimensionless form of radial equation is expressed as ˆ l Rn ,l = 2 Enr ,l Rn ,l H r r ¯h ω

(2284)

where the dimensionless effective Hamiltonian for angular momentum l is given by   ˆl = H Aˆ†l Aˆl + 2 ( l + 1 ) (2285) and the effective Hamiltonian for angular momentum l + 1 is given by   † ˆ ˆ ˆ Hl+1 = Al Al + 2 l

(2286)

By using the radial equation in the form ˆ l Rn ,l H r   Aˆ†l Aˆl + 2 ( l + 1 ) Rnr ,l and then on pre-multiplying by Aˆl , one finds   Aˆl Aˆ†l Aˆl + 2 ( l + 1 ) Rnr ,l   Aˆl Aˆ†l + 2 ( l + 1 ) Aˆl Rnr ,l   ˆ l+1 + 2 Aˆl Rn ,l H r 44 In

= =

Enr ,l Rnr ,l ¯h ω Enr ,l 2 Rnr ,l ¯h ω 2

Enr ,l ˆ Al Rnr ,l ¯h ω Enr ,l ˆ = 2 Al Rnr ,l ¯h ω Enr ,l ˆ = 2 Al Rnr ,l ¯h ω =

(2287)

2

(2288)

d dimensions, the commutator of the raising and lowering operators is evaluated ˆ† , A ˆl ] = − 2 − 2 l + d − 1 [A l ρ2

and the product of the operators is given by 2 d−1 ∂ l(l + d − 2) ˆ† A ˆl = − ∂ A − + + ρ2 − ( 2 l + d ) l ∂ρ2 ρ ∂ρ ρ2

(2282)

478

ˆ l+1 since Hence, we have shown that Aˆl Rnr ,l is an eigenstate of H   Enr ,l ˆ l+1 Aˆl Rn ,l = 2 H − 1 Aˆl Rnr ,l r ¯h ω

(2289)

Therefore, the lowering operator Aˆl acting on Rnr ,l produces another energy eigenfunction with indices (nr − 1, l + 1) Aˆl Rnr ,l ∝ Rnr −1,l+1

(2290)

and with energy eigenvalue given by Enr −1,l+1 = Enr ,l − ¯h ω

(2291)

Since the energy is bounded from below there is a minimum value of nr (defined as nr = 0), so that the lowering operator acting on this state with nr = 0 vanishes Aˆl R0,l = 0 (2292) On using this condition in eqn(2287), one finds the energy eigenvalue equation simplifies to E0,l ( l + 1 ) R0,l = R0,l (2293) ¯h ω Hence, the energy eigenvalues for the states with nr = 0 are given by   E0,l = h ¯ ω l + 1 (2294) The radial eigenfunctions are found from the condition



Aˆl R0,l  ∂ l R0,l + ρ − ∂ρ ρ

which can be re-expressed as 

∂R0,l ∂ρ

R0,l

= 0 = 0

(2295)

l ρ

(2296)

 = −ρ +

This can be integrated to yield l

R0,l = C ρ exp



ρ2 − 2

 (2297)

The above form of the energy eigenfunction is also found for the isotropic ddimensional harmonic oscillator, and the energy eigenvalue corresponding to this eigenfunction is given by   d E0,l = h ¯ ω l + (2298) 2 479

The radial eigenfunctions for larger values of nr can be obtained from the action of the raising operators. This can be seen by considering the radial equation with angular momentum l + 1 ˆ l+1 Rn ,l+1 H r   Aˆl Aˆ†l + 2 l Rnr ,l+1 and then pre-multiplying by the   Aˆ†l Aˆl Aˆ†l + 2 l   Aˆ†l Aˆl + 2 l Aˆ†l   ˆ Hl − 2 Aˆ†l

Enr ,l+1 Rnr ,l+1 ¯h ω Enr ,l+1 = 2 Rnr ,l+1 ¯h ω

=

2

(2299)

raising operator Aˆ†l Rnr ,l+1 Rnr ,l+1 Rnr ,l+1

Enr ,l+1 ¯h ω Enr ,l+1 = 2 ¯h ω Enr ,l+1 = 2 ¯h ω =

2

Aˆ†l Rnr ,l+1 Aˆ†l Rnr ,l+1 Aˆ†l Rnr ,l+1

(2300)

Hence, we have ˆ l Aˆ† Rn ,l+1 = 2 H r l



Enr ,l+1 +1 ¯h ω



Aˆ†l Rnr ,l+1

(2301)

so the raising operator acting on a radial eigenfunction with quantum numbers (nr , l + 1) produces an eigenstate with quantum numbers (nr + 1, l) and eigenvalue Enr +1,l = Enr ,l+1 + ¯h ω (2302) Therefore, the eigenfunctions with higher values of n are found from Rnr +1,l ∝ Aˆ†l Rnr ,l+1 The energy eigenvalues are found as   Enr ,l = h ¯ ω 2 nr + | l | + 1

(2303)

(2304)

This completely solves for the energy spectrum and the energy eigenfunctions when Ω = 0. The classical states can be recovered as coherent states, since the ratios of the rotational frequencies to the vibrational frequencies are rational45 . In the more general case, one finds   Ω Enr ,l = h ¯ ω 2 nr + | l | + l + 1 ω 45 R.

Balian and C. Bloch, Ann. Phys. N.Y. 69, 76 (1972).

480

(2305)

In the general case, the ratios of the frequencies are not rational. The energy spectrum can be plotted as a function of the de-tuning parameter ν defined by   Ω − ω ν = (2306) 2ω The spectrum is shown in fig(121). For negative rational values of ν the spec-

Figure 121: The energy spectrum for the rotated planar harmonic oscillator as a function of the de-tuning parameter ν. trum shows gaps46 . These gaps are related to the occurrence of periodic orbits. The gaps are easily seen in the higher energy portions of the spectrum, shown in fig(122). For ν = 0 a catastrophe occurs, all the energy levels become infinitely degenerate and each group of levels is separated by the energy difference 2 ¯h ω.

46 R.K.

Bhaduhri, S. Li, K. Tanaka and J.D. Waddington, J. Phys. A 27, L663 (1994).

481

Figure 122: The higher-energy eigenvalues for the rotated planar harmonic oscillator as a function of the de-tuning.

6

Dirac Formulation

Schr¨ odinger introduced wave mechanics in which a quantum state was described in terms of a wave function and physical measurements were described by differential operators. On the other hand, Heisenberg introduced matrix mechanics in which states were represented by column vectors, similar to the column vectors used in our discussion of spin, and measurements were treated as matrices similar to the spin matrices. These two formulations of quantum mechanics are identical in physical content. The equivalence was first seen by Dirac and this led to his abstract formulation.

6.1

Dirac Notation

Dirac introduced an abstract notation for states and their duals, and also for the operators.

482

6.1.1

Bracket Notation

In the Dirac formulation, physical states are represented by the so called kets, which are written as |Ψ > (2307) These kets obey the principle of linear superposition X |Ψ > = Cn | φn >

(2308)

n

where Cn are complex numbers. Generally, if the discrete index n is supplemented by a continuous index47 denoted by ξn , the superposition can be expressed in terms of a sum and an integral Z X |Ψ > = Cn | φn > + dξ Cξ | φξ > (2309) n

Dirac also introduced a dual space of bras. The bras are denoted as < Ψ | and are defined as the duals to the kets | Ψ >. They are the mirror image of the ket states. More precisely, they are defined by the scalar products with the kets defined as a complex number given by   < Φ| |Ψ > = < Φ|Ψ > (2310) The scalar product is linear, so that   < Φ | C1 | Ψ1 > + C2 | Ψ2 > = C1 < Φ | Ψ1 > + C2 < Φ | Ψ2 > (2311) A bra vector is completely defined when the scalar product with every ket is known. Thus, if the scalar product with every ket vector is zero, then the bra vector is also zero. < Φ|Ψ > = 0

for all

|Ψ >

(2312)

then < Φ| = 0

(2313)

We assume that there is a one to one correspondence between the bras and the kets. That is, the bra corresponding to the sum of the kets | φ1 > + | φ2 > 47 For example, if the set of states | φ n > corresponds to the eigenstates of some operator, the eigenvalue may have both discrete and continuous portions of its spectrum.

483

is the sum of the equivalent bras < φ1 | + < φ2 |. Also, the bra corresponding to the ket C | φ > is given by C ∗ < φ |, which involves complex conjugation. Because of this assumption, every physical state can also be represented by a bra as much as it can be represented by a ket. The scalar product is defined such that the complex number equal to the scalar product < Φ|Ψ >

(2314)

is related to the number given by the scalar product of the dual vectors < Ψ|Φ >

(2315)

via complex conjugation. That is, we require that the scalar product satisfies < Φ | Ψ > = < Ψ | Φ >∗

(2316)

In particular, the normalization of a state can be defined as the scalar product with itself, < Ψ | Ψ > = < Ψ | Ψ >∗

(2317)

which must be real. We shall require that the scalar product be defined such that the normalization is both real number and positive < Ψ|Ψ > ≥ 0

(2318)

Two states | Φ > and | Ψ > are defined to be orthogonal if their scalar product is zero < Φ|Ψ > = 0 (2319)

6.1.2

Operators

A physical measurement is described by an operator Aˆ which is defined by its effect on a ket, and in general, transforms a ket to another ket, Aˆ | Ψ > = | Φ >

484

(2320)

The operators are linear in that its effect on a linear superposition is expressed as a linear combination on the individual states X Aˆ | Ψ > = Cn Aˆ | φn > (2321) n

The effect of an operator on a bra can be defined via the scalar product   < Φ | Aˆ | Ψ > = < Φ | Aˆ | Ψ >   = < Φ | Aˆ | Ψ > (2322) This defines the effect of the operator Aˆ on the bra < Φ |, < Φ | Aˆ

6.1.3

(2323)

Adjoints and Hermitean Operators

The Hermitean conjugate of an operator Aˆ is defined as the operator Aˆ† such that the bra < Ψ | Aˆ (2324) is equivalent to the ket Aˆ† | Ψ >

(2325)

for any state | Ψ >. In particular, this means that < Ψ | Aˆ | Φ > = < Φ | Aˆ† | Ψ >∗

(2326)

Physical operators are represented by Hermitean operators such that Aˆ† = Aˆ

(2327)

The eigenvalue equation of an operator Aˆ is Aˆ | φn > = an | φn >

(2328)

where an are the eigenvalues and | φn > are the eigenstates. It can be proved that the eigenstates of a Hermitean operator form a complete orthogonal set and their eigenvalues are real.

485

An arbitrary state can be represented as a linear superposition of a complete orthonormal set of states | φn > . That is, an arbitrary state | Ψ > is re-written in terms of the expansion X |Ψ > = Cn | φn > (2329) n

where, due to orthonormality of the complete set, the expansion coefficient is evaluated as Cn = < φn | Ψ > (2330) Thus, we have |Ψ > =

X

< φn | Ψ > | φn >

(2331)

| φn > < φn | Ψ >

(2332)

n

which can be re-arranged as |Ψ > =

X n

Since | Ψ > is completely arbitrary, the relation is independent of | Ψ > . Omitting the arbitrary state leads to the completeness relation, in which the identity operator Iˆ is written as X Iˆ = | φn > < φn | (2333) n

6.1.4

Representation of Operators

An arbitrary operator Aˆ can be represented in much the same way. We define the operator via its effect on an arbitrary state | Ψ > Aˆ | Ψ > = | Φ >

(2334)

Then, with the analogous expression for | Φ > with expansion coefficients Bm , we find that the effect of the operator is given by X X Aˆ Cn | φn > = Bm | φm > (2335) n

m

Taking the scalar product of the above equation with the bra < φm | , one finds that the matrix elements of the operator < φm | Aˆ | φn > satisfy the equation X < φm | Aˆ | Ψ > = < φm | Aˆ | φn > Cn n

= Bm 486

(2336)

Inserting the expression for Cn , one finds that the effect of an operator on an arbitrary state | Ψ > can be expressed directly in terms of the matrix elements  X  ˆ ˆ < φm | A | Ψ > = < φm | A | φn > < φn | Ψ > n

= Bm

(2337)

Hence, we can use the expression for Bm in the equation defining the effect of the operator X Aˆ | Ψ > = Bm | φm > m

=

X

 | φm >

< φm | Aˆ | φn >

 < φn | Ψ >

m,n

(2338) Thus, as | Ψ > is arbitrary the above equation is independent of the choice of | Ψ > , so we may omit it. Hence, the operator Aˆ can be decomposed as   X Aˆ = | φm > < φm | Aˆ | φn > < φn | (2339) m,n

The operator is completely defined by its matrix elements between a complete set of states.

6.2

Representations

The position space states | r > are eigenstates of the Hermitean position operator ˆr, ˆr | r > = r | r > (2340) Due to the Hermiticity, the eigenvalues of ˆr are real. The orthogonality of the eigenstates are expressed as < r | r0 > = δ 3 ( r − r0 )

(2341)

ˆ The completeness condition is expressed in terms of the identity operator I, where Z (2342) Iˆ = d3 r | r > < r |

The position space representation associates a complex number Ψ(r) (the wave function) with every scalar product of a state represented by a ket | Ψ > and the bra representing a position state < r | < r | Ψ > = Ψ(r) 487

(2343)

alternatively, one obtains the complex conjugate of the wave function from < Ψ | r > = Ψ∗ (r)

(2344)

A general scalar product can be evaluated using the completeness relation for the position states, which involves the identity operator Z Iˆ = d3 r | r > < r | (2345) as Z < Φ|Ψ >

= Z =

d3 r < Φ | r > < r | Ψ > d3 r Φ∗ (r) Ψ(r)

(2346)

Thus, the scalar product has a position space representation as the overlap operator. The position space representation of an operator Aˆ can also be found from the matrix elements between two arbitrary states, by inserting complete sets of states Z < Φ | Aˆ | Ψ > = d3 r < Φ | r > < r | Aˆ | Ψ > Z Z = d3 r d3 r0 < Φ | r > < r | Aˆ | r0 > < r0 | Ψ > (2347) The position space representation of the operator Aˆ is given by < r | Aˆ | r0 >

(2348)

For a local operator, the operator only depends on the position variable and the derivative at one point in space, so we have < r | Aˆ | r0 > = δ 3 ( r − r0 ) A( r0 ; − i ¯h ∇0 )

(2349)

Thus, on integrating over r one obtains the usual expression Z ˆ < Φ|A|Ψ > = d3 r0 < Φ | r0 > A( r0 ; − i ¯h ∇0 ) < r0 | Ψ > Z = d3 r Φ∗ (r) A( r ; − i ¯h ∇ ) Ψ(r) (2350) In this expression, the differential operator A( r ; − i ¯h ∇ ) is that which appears in the usual formulation of Schr¨odinger’s wave mechanics.

488

6.3

Gram-Schmidt Orthogonalization

Given a complete set of normalized states | φm > , that are not orthogonal, one can construct an orthonormal set | Ψn > via the process of Gram-Schmidt orthogonalization. This is achieved by taking the first state as | Ψ0 > = | φ0 >

(2351)

which is normalized to unity. The second state in our orthogonal set is constructed as   | Ψ1 > = N1 | φ1 > − | Ψ0 > < Ψ0 | φ1 > (2352) This second state is orthogonal to | Ψ0 > as can be seen by forming the scalar product, and noting that the first state is normalized to unity. The normalization of the second state N1 is chosen such that this state is also normalized to unity. The higher order states are found by orthogonalizing to the properly normalized states found earlier, i.e.,  | Ψn > = Nn

| φn > −

m=n−1 X

 | Ψm > < Ψm | φn >

(2353)

m=0

and finding the normalization Nn constants before proceeding to find the next state. The normalization constants are given by Nn−2 = 1 −

m=n−1 X

| < Ψm | φn > |2

(2354)

m=0

Hence, given any complete set of states, it is possible to transform them into an orthonormal set.

489

7

Appendices

A: Non-spreading probability densities The spreading of a wave packet (such as a Gaussian) is a feature which is often found in quantum mechanics. Normalizable wave functions are interpretable in terms of the motion of a single particle. Non-normalizable wave functions should be thought of describing an ensemble of particles. There are only two non-spreading wave functions, the plane wave which corresponds to a uniform beam of particles with momentum h ¯ k, and the Airy wave function which we shall discuss below. The Airy wave function is given at t = 0 by   Bx Ψ(x; 0) = Ai 2 h3

(2355)

which has the Fourier transform 2

Φ(k; 0) = √

1 h ¯3 exp 2π B

 i

¯ 2 k3 h 3 B3

 (2356)

Hence, in momentum representation, the time-dependent wave function for free particles is given by 2

Φ(k; t) = √

1 h ¯3 exp 2π B



 i

¯h2 k 3 ¯h k 2 − t 3 B3 2m

 (2357)

Therefore, we find that the real-space wave function is given by the inverse Fourier transform    2 Z ∞ 1 ¯ h3 ¯h2 k 3 ¯h k 2 Ψ(x; t) = dk exp i k x + − t (2358) 2 π B −∞ 3 B3 2m On changing the variable of integration from k to k 0 k → k0 = k −

B3 t 2 m ¯h

(2359)

one finds that the integration results in another Airy function. The result is       B 3 t2 B B3 t B 3 t2 x − Ψ(x; t) = Ai exp i x − (2360) 2 4 m2 2 m ¯h 6 m2 h3 ¯ Therefore, the probability density is given by    B B 3 t2 | Ψ(x; t) |2 = Ai2 x − 2 4 m2 ¯h 3

490

(2361)

Probability density for the Airy wave packet 2

|Ψ(x;0)|2

1.5

1

0.5

0 -12

-8

-4

0

4

x

Figure 123: The initial probability density |Ψ(x; 0)|2 for an Airy wavepacket. The wavepacket does nor disperse with increasing t, but it does accelerate. which propagates without distortion and has an uniform acceleration given by B3 2 m2

(2362)

This peculiar result can be understood on the basis of a comparison of the semi-classical form of the wave function with a family of classical trajectories48 . The semi-classical limit of the Airy function is given by   14   3  2 1 h3 ¯ 2 −Bx 2 π Ψ(x; 0) ∼ √ sin + (2363) 2 3 4 π −Bx ¯h 3 The semi-classical wave function is interpreted as a superposition of forward and backward travelling particles with actions S± given by   32 2 S± = ± − Bx (2364) 3 From Hamilton-Jacobi theory, the momenta are given by   12 ∂S± p± = = ± − Bx ∂x 48 M.V.

Berry and N.L. Balazs, Am. J. Phys. 47, 264 (1979).

491

(2365)

This implies that there is a phase-space relationship given by p2 (2366) B3 The Airy function can be considered as resembling an infinite ensemble of classical particles, with initial positions given by X0 (p) and each particle is parameterized by its momentum p. These particles map out straight-line trajectories given by p t (2367) X(p; t) = X0 (p) + m The envelope of the family of trajectories is a parabola. This can be seen by X0 (p) = −

Envelope of a family of trajectories 6 4

x(t)=B3 t2/(4 m2)

x

2 0 -2

X(p,t)=-p2/B3+pt/m

-4 -6 -5

-4

-3

-2

-1

0

1

2

3

4

5

t

Figure 124: The family of trajectories represented by an Airy wave packet, and its parabolic envelope. noting that at fixed value of t, the maximum value of X(p, t) corresponds to the value pt determined from ∂X(p, t) = 0 (2368) ∂p pt which yields the relation B3 t (2369) 2m Since the family of trajectories touch the envelope at pt , the envelope x(t) is given by p2 pt x(t) = X(pt , t) = − t3 + t (2370) B m pt =

492

or

B3 2 B2 2 B3 2 t + t = t (2371) 2 2 4m 2m 4 m2 Hence the envelope is parabolic. This parabolic envelope represents the boundaries of classical motion and closely follows the motion of the maximum value of the quantum mechanical probability density. x(t) = −

—————————————————————————————————— B: Bound States within the Continuum von Neumann and Wigner found that it is possible to construct potentials which have bound states at energies that are higher than the lowest energy of the continuum spectra49 . However, in these unusual examples, the potential contains infinitely many oscillations. The method used is based on the solution of the field-free energy eigenvalue equation, with any fixed values of the angular momentum and energy. The energy eigenvalue equation for a free particle with angular momentum l has the form   1 ∂ l(l + 1) 2 ∂Rl r + Rl = k 2 Rl (2372) − 2 r ∂r ∂r r2 where the energy eigenvalue E is positive and is related to k via ¯h2 k 2 > 0 2m The radial equation has the non-normalizable solution E =

Rl (r) = jl (kr)

(2373)

(2374)

corresponding to the spherical Bessel function. This energy eigenstate is extended and corresponds to the continuous portion of the energy spectrum. von Neumann and Wigner suggested that any continuum state may be modified so as to produce a localized or normalizable state, by multiplying by some 1 function f (r) which vanishes faster than r− 2 as r → ∞. As they showed, it is possible to construct a potential such that the modified state is an eigenstate of the Hamiltonian with the same value of the energy. Extending these results to finite values of l, Stillinger and Herrick50 expressed the radial function as Rl (r) = jl (kr) f (r)

(2375)

Since this wave function is required to satisfy the energy eigenvalue equation with fixed quantum numbers l and E    1 ∂ ∂Rl l(l + 1) 2m − 2 V (r) Rl = k 2 Rl (2376) r2 + r ∂r ∂r r2 ¯h2 49 J.

von Neumann and E. Wigner, Phys. Z. 30, 465 (1929). Stillinger and D.R. Herrick, Phys. Rev. A, 11 446 (1976).

50 F.H.

493

the potential is specified in terms of derivatives of the eigenfunction  0  0 2m 1 f (r) f ”(r) jl (kr) + + 2 V (r) = 2 k j (kr) k r f (r) f (r) ¯h l

(2377)

0

(r) The potential falls to zero at infinity, if both ff (r) and ff”(r) (r) vanish as r → ∞. For an arbitrary decaying form of f (r), the potential would have poles at the zeros of jl (kr). If the potential is required to be finite, then the poles 0 (r) , originating from the zeros of jl (kr) should be suppressed by the zeros of ff (r) and furthermore f (r) should have no zeros. This suggests that the function f (r) should be an analytic function of the variable s(r), where Z r s(r) = k dr0 ( kr0 )2l+3 jl2 (kr0 ) 0   1 2l+4 2 2 ( kr ) jl (kr) + jl+1 (kr) (2378) = 4(l + 1)

This has the effect that derivative of s(r) vanishes at the zeros of jl (kr), but s(r) never decreases. The function f (r) may be chosen as f (r) =

A2

1 + s(r)

(2379)

so that f (r) vanish at infinity and Rl (r) is normalizable. Due to the multiplicative factor of jl (kr) the wave function has an infinite number of nodes. The potential is then found as   ( kr )4l+6 jl4 (kr) 2 ( kr )2l+3 jl (kr) jl0 (kr) + ( l + 52 ) ( kr )2l+2 jl2 (kr) V (r) = k 2 − ( A2 + s(r) )2 A2 + s(r) (2380) The long-ranged behavior of the potential is then found to be given by V (r) ∼ ( − 1 )l 8 ( l + 1 ) k 2

sin 2 kr 2 kr

(2381)

The absence of an outgoing “spherical” wave at infinity can be thought of as being caused by the infinite sequence of reflections by subsequent maxima in the potential. At each maximum, there is always some reflection, even though the particle may have an energy greater than the potential barriers. Indeed, the energy of the bound state may even exceed the maximum value of the potential. In this truly exceptional case, even a classical particle would never be localized. —————————————————————————————————— C: The Symmetric Rotor We shall consider the quantum mechanics describing the rotation of a rigid body. Initially, the body is assumed to have non-identical moments of inertia 494

Radial wave function for l = 0 0.3

R0(r)

0.2

0.1

0 0

0.5

1

1.5

2

-0.1

kr/π

Figure 125: The radial wave function R0 (r) for an l = 0 bound state with positive energy considered by von Neumann and Wigner. I1 , I2 and I3 . We shall describe the orientation of the rigid body by the Euler angles (α, β, γ) which describe the orientation of a set of axes embedded in the body with respect to an external fixed frame of reference. The Euler angles are angles which describe three successive rotations of the body, which bring it from a configuration in which the axes embedded in the body coincide with the axes of the fixed reference frame. The first angle α describes the first rotation of the body about the z-axis. This rotation rotates the x and y axes embedded in the body into new positions, x0 and y 0 . The second rotation usually is prescribed as a rotation about the y 0 -axis, through the angle β. The second rotation brings the z-axis into the position z 00 . The z 00 -axis has polar coordinates (β, α). The last rotation is a rotation through an angle γ about the z 00 -axis. The rotation can be represented by three by three matrices which acts on the Cartesian coordinates (x1 , x2 , x3 ) of a vector. Under the rotation, the point r rigidly attached to the body is transformed to the point r0 , by r0

ˆ = R(α, β, γ) r ˆ ˆ y0 (β) R ˆ z (α) r 00 = Rz (γ) R

(2382)

The rotation through the angle α about the z-axis is represented by the three 495

Oscillating Potential

V(r) [units of h2k2/m]

6 4 2

E 0 -2 -4 0

0.5

1

1.5

2

kr/π

Figure 126: The oscillating potential V (r) of von Neumann and Wigner that produces an l = 0 bound state with positive energy. by three matrix 

cos α ˆ z (α) =  sin α R 0

− sin α cos α 0

 0 0  1

(2383)

The rotation through an angle β about the y 0 -axis is described by ˆ y0 (β) = R ˆ z (α) R ˆ y (β) R ˆ z (α)−1 R where



cos β ˆ y (β) =  0 R − sin β

0 1 0

 sin β 0  cos β

(2384)

(2385)

The final rotation is the rotation about the angle γ about the z 00 axis. This is represented by ˆ z00 (γ) R

where

ˆ y0 (β) R ˆ z (α) R ˆ z (γ) R ˆ z (α)−1 R ˆ y0 (β)−1 = R ˆ z (α) R ˆ y (β) R ˆ z (γ) R ˆ y (β)−1 R ˆ z (α)−1 = R 

cos γ ˆ z (γ) =  sin γ R 0

496

− sin γ cos γ 0

 0 0  1

(2386)

(2387)

Hence, the combined rotation is given by the product of the rotations about the fixed axes as ˆ ˆ z (α) R ˆ y (β) R ˆ z (γ) R(α, β, γ) = R (2388) and has the representation     cos α − sin α 0 cos β cos γ − cos β sin γ sin β ˆ cos α 0   sin γ cos γ 0  R(α, β, γ) =  sin α − sin β cos γ sin β sin γ cos β 0 0 1   cos α cos β cos γ − sin α sin γ − cos α cos β sin γ − sin α cos γ cos α sin β =  sin α cos β cos γ + cos α sin γ − sin α cos β sin γ + cos α cos γ sin α sin β  − sin β cos γ sin β sin γ cos β The angle of rotation ψ can be found by evaluating the trace. The trace is found as cos β + cos(α + γ) ( 1 + cos β ) (2389) which is equal to 1 + 2 cos ψ. Hence, cos ψ = cos2

β β cos(α + γ) − sin2 2 2

(2390)

The kinetic energy of the symmetric rotor is described by the Lagrangian   1 2 2 2 L = I3 ω3 + I2 ω2 + I1 ω1 (2391) 2 where ω1 , ω2 and ω3 are the components of the angular velocity of rotation about the three body-fixed principal axes. We shall first express this Lagrangian in terms of the Euler angles and then determine the Hamiltonian. The angular velocity vector ω can be expressed as ω = α˙ eˆα + β˙ eˆβ + γ˙ eˆγ

(2392)

The unit vectors eˆα , eˆβ , eˆγ are to be expressed in terms of the unit vectors in the body-fixed coordinate system. Since the angular velocity γ˙ takes place around the body fixed z-axis (z 00 ), one has eˆγ = eˆ3

(2393)

The angular velocity β˙ occurs around the y 0 -axis. The y 0 -axis is in the eˆ1 − eˆ2 plane, since the rotation γ is around eˆ3 and keeps the eˆ1 − eˆ2 plane invariant. The azimuthal angle of the y 0 axis is π2 − γ. Hence eˆβ

π π − γ) + eˆ2 sin( − γ) 2 2 = eˆ1 sin γ + eˆ2 cos γ

= eˆ1 cos(

497

(2394)

The final component of the rotation is around the α axis, which is aligned along the original z-axis. The eˆα axis has polar angle β and azimuthal angle π − γ relative to the fixed body axes.   + eˆ3 cos β (2395) eˆα = sin β − cos γ eˆ1 + sin γ eˆ2 Hence, the components of the angular velocity with respect to the body-fixed axes are expressed in terms of the Euler angles via ω1 ω2 ω3

= − sin β cos γ α˙ + sin γ β˙ = sin β sin γ α˙ + cos γ β˙ = cos β α˙ + γ˙

(2396)

Hence, the Lagrangian is given as   2  2  2  1 ˙ ˙ L = I1 − sin β cos γ α˙ +sin γ β + I2 sin β sin γ α˙ + cos γ β + I3 cos β α˙ + γ˙ 2 (2397) For the symmetric top, one has I1 = I2 , therefore, the Lagrangian simplifies to     2  1 2 2 2 ˙ L = I1 sin β α˙ + β + I3 cos β α˙ + γ˙ (2398) 2 The momenta are given by pα

=

∂L ∂ α˙ 

2

= I1 sin β α˙ + I3 cos β and pβ =

 cos β α˙ + γ˙

∂L = I1 β˙ ∂ β˙

(2399)

(2400)

and finally pγ =

∂L = I3 ∂ β˙



 cos β α˙ + γ˙

(2401)

The Hamiltonian H is given by H

= pα α˙ + pβ β˙ + pγ γ˙ − L  2 1 1 1 p − cos β p + = p2β + p2 (2402) α γ 2 2 I1 2 I3 γ 2 I1 sin β

This Hamiltonian is quantized as ˆ H

=

1 1 pˆ2 + 2 I1 β 2 I1 sin2 β

2



498

pˆα − cos β pˆγ

+

1 pˆ2 (2403) 2 I3 γ

It is seen that pˆα and pˆγ commute with the Hamiltonian and, therefore, one can ˆ The simultaneous eigenfunctions find simultaneous eigenstates of pˆα , pˆγ and H. are written in the form     i 1 Ψ(α, β, γ) = exp ( α pα + γ pγ ) φ(β) (2404) 2π ¯h Since the wave functions are required to be single-valued, the quantum numbers are determined to be pα = mα ¯h and pγ = mγ ¯h where (mα , mγ ) are either both integers or they are both half-integers. On substituting the above form, the expression for the eigenvalue equation reduces to  2     1 2 1 1 p − cos β p φ(β) = E − p φ(β) (2405) pˆ2β + α γ 2 I1 2 I3 γ sin2 β Since β is a polar angle, the above equation takes the explicit form    2   ¯h2 m2γ h2 ¯ 1 ∂ ∂φ ¯h2 − sin β + mα − cos β mγ φ(β) = E − φ(β) 2 I1 sin β ∂β ∂β 2 I3 2 I1 sin2 β (2406) or    2   ∂ ∂φ I1 2 2 I1 E 2 − sin β sin β + mα − cos β mγ + sin β m − φ(β) = 0 ∂β ∂β I3 γ ¯h2 (2407) Clearly, this can be rewritten as an eigenvalue equation in terms of the variable z = cos β, where z is restricted to the range 1 ≥ z ≥ − 1. Therefore, we have     ∂ 2 2 ∂φ(β) 2 2 2 (1−z ) + mα + mγ − λ − 2 mα mγ z + λ z φ(β) = 0 −(1−z ) ∂z ∂z (2408) where λ is related to the energy eigenvalue via   I1 2 I1 E + 1− m2γ (2409) λ = I3 ¯h2 The eigenvalue equation has regular singular points at z = ± 1. Near these points, the solution has a non-analytic form, and is expressed as φ(β) = ( 1 − z )µ ( 1 + z )ν F (z)

(2410)

where F (z) is an analytic function, and the indices µ and ν are to be determined. The indices are determined by considering the most singular terms, and are found to be given by   mα − mγ µ = ± (2411) 2 and

 ν = ±

mα + mγ 2

499

 (2412)

The solutions with the negative values for the exponents are discarded since the solution is required to be finite and normalizable. The function F (z) then must satisfy an equation of the form     ∂2F ∂F + 2 ( 1 + m ) z − m + m ( m + 1 ) − λ F = 0 − ( 1 − z2 ) α γ α α ∂z 2 ∂z (2413) For convenience, we have assumed that mα and mγ are both positive and that mα > mγ . The eigenvalue λ is found by using the Fr¨obenius method, expanding in powers about a singular point and requiring that the series expansion terminates, yielding a polynomial for F (z). This requires that r 1 1 ( mα + ) ± + λ (2414) 2 4 is a positive integer. Hence, we find that λ = j(j + 1)

(2415)

and the energy eigenvalue is given by E =

¯2 h ¯h2 m2α ¯ 2 m2α h j (j + 1 ) + − 2 I1 2 I3 2 I1

(2416)

as could have been expected. —————————————————————————————————— c “ 2005 by Peter S. Riseborough”

500

Smile Life

When life gives you a hundred reasons to cry, show life that you have a thousand reasons to smile

Get in touch

© Copyright 2015 - 2024 PDFFOX.COM - All rights reserved.